Actin: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
English Fig (talk | contribs)
inserted translated text ~~~~
Line 1: Line 1:
[[Image:Actin with ADP highlighted.png|thumb|G-actin ([[Protein Data Bank|PDB]] code: [http://www.pdbe.org/1j6z 1j6z]). [[adenosine diphosphate|ADP]] and the divalent cation are highlighted.]]
[[Image:Actin with ADP highlighted.png|thumb|G-actin ([[Protein Data Bank|PDB]] code: [http://www.pdbe.org/1j6z 1j6z]). [[adenosine diphosphate|ADP]] and the divalent cation are highlighted.]]
[[Image:Actin filament atomic model.png|thumb|F-actin; surface representation of 13 subunit repeat based on Ken Holmes' actin filament model]]

{{Infobox protein family
{{Infobox protein family
| Symbol = Actin
| Symbol =Actin
| Name = Actin
| Name ={{PAGENAME}}
| Image =Actin_with_ADP_highlighted.png
| Pfam = PF00022
| Width =250px
| Pfam_clan = CL0108
| image_caption =G-actin ([[Protein Data Bank|PDB]] code: [http://www.pdbe.org/1j6z 1j6z]). The [[adenosine diphosphate|ADP]] molecule’s [[active site]] and divalent cation are highlighted.<ref>[http://www.pdb.org/pdb/explore/explore.do?structureId=1J6Z Uncomplexed Actin], [[Protein Data Bank]]</ref>
| InterPro = IPR004000
| SMART =
| Pfam =PF00022
| InterPro =IPR004000 | SMART =
| PROSITE = PDOC00340
| PROSITE =PDOC00340
| MEROPS =
| SCOP = 2btf
| SCOP =2btf
| TCDB =
| TCDB =
| OPM family =
| OPM family =
| OPM protein =
| OPM protein =
| PDB = {{PDB|link=no|1atn}} actin from ''[[Oryctolagus cuniculus]]''
| CAZy =
| CDD =
}}
}}


'''Actin''' is a [[Globular protein|globular]], roughly 42-[[kDa]] multi-functional protein found in all [[Eukaryote|eukaryotic cells]] (the only known exception being [[nematode]] sperm), where it may be present at concentrations of over 100 [[Micromolar|μM]]. It is also one of the most highly-[[Conserved sequence|conserved]] proteins, differing by no more than 20% in [[species]] as diverse as [[algae]] and [[human]]s. Actin is the [[monomer]]ic subunit of two types of filaments in cells: [[microfilaments]], one of the three major components of the [[cytoskeleton]], and thin filaments, part of the contractile apparatus in muscle cells. Thus, actin participates in many important cellular processes, including [[#Actomyosin filaments|muscle contraction]], cell [[motility]], cell division and [[cytokinesis]], vesicle and organelle movement, [[cell signaling]], and the establishment and maintenance of [[cell junction]]s and cell shape. Many of these processes are mediated by extensive and intimate interactions of actin with cellular membranes.<ref>{{cite journal |author=Doherty GJ and McMahon HT |title=Mediation, Modulation and Consequences of Membrane-Cytoskeleton Interactions |journal=Annual Review of Biophysics |volume=37 |pages=65–95 |year=2008 |pmid=18573073 |url=http://arjournals.annualreviews.org/doi/abs/10.1146/annurev.biophys.37.032807.125912 |doi=10.1146/annurev.biophys.37.032807.125912}}</ref> In vertebrates, three main groups of actin [[isoforms]], [[ACTA1|alpha]], [[ACTB|beta]], and [[ACTG1|gamma]] have been identified. The alpha actins, found in muscle tissues, are a major constituent of the contractile apparatus. The beta and gamma actins coexist in most cell types as components of the [[cytoskeleton]], and as [[chemical mediator|mediators]] of internal cell motility.


==Formation of thin filament==
[[Image:Thin filament formation.svg|thumb|500px|center|Formation of thin filament]]


[[Image:Actin filament atomic model.png|thumb|270px|F-actin; surface representation of a repetition of 13 subunits based on Ken Holmes' actin filament model.<ref name=Holmes1990>{{cite journal | surname1= Holmes | name1= K.C. | surname2= Popp | name2= D. | surname3= Gebhard | name3= W. | surname4= Kabsch | name4= W. | year= 1990 | title = Atomic model of the actin filament | pub-journal= Nature | volume= 347 | number= 6288 | page = 44–49 | doi = 10.1038/347044a0 | url = http://www.nature.com/nature/journal/v347/n6288/abs/347044a0.html | format= w}}</ref>]]
==Genetics==

[[Image:Adherens Junctions structural proteins.svg|thumb|350px|]]Principal interactions of structural proteins are at [[cadherin]]-based adherens junctions. Actin filaments are linked to α-[[actinin]] and to the membrane through [[vinculin]]. The head domain of vinculin associates to E-cadherin via α-, β-, and γ-catenins. The tail domain of vinculin binds to membrane lipids and to actin filaments.
'''Actin''' is a [[Globular protein|globular]] multi-functional protein that forms [[microfilament]]s. It is found in all [[Eukaryote|eukaryotic cells]] (the only known exception being [[nematode]] sperm), where it may be present at concentrations of over 100 [[Micromolar|μM]]. Actin is roughly 42-[[kDa]] in size and it is the [[monomer]]ic subunit of two types of filaments in cells: [[microfilaments]], one of the three major components of the [[cytoskeleton]], and thin filaments, part of the contractile apparatus in muscle cells. It can be present as either a free [[monomer]] called '''G-actin''' or as part of a linear [[polymer]] '''microfilament''' called '''F-actin''' both of which are essential for such important cellular functions as the mobility and contraction of [[cell (biology)|cell]]s during [[cell division]].
The protein actin is one of the most highly conserved throughout evolution because it interacts with a large number of other proteins, with 80.2% sequence [[Conservation (genetics)|conservation]] at the [[gene]] level between ''[[Human|Homo sapiens]]'' and ''[[Saccharomyces cerevisiae]]'' (a species of yeast), and 95% conservation of the [[primary structure]] of the protein product.

Thus, actin participates in many important cellular processes, including [[#Outline of a muscle contraction|muscle contraction]], cell [[motility]], cell division and [[cytokinesis]], vesicle and organelle movement, [[cell signalling]], and the establishment and maintenance of [[cell junction]]s and cell shape. Many of these processes are mediated by extensive and intimate interactions of actin with cellular membranes.<ref>{{cite journal |author=Doherty GJ and McMahon HT |title=Mediation, Modulation and Consequences of Membrane-Cytoskeleton Interactions |journal=Annual Review of Biophysics |volume=37 |pages=65–95 |year=2008 |pmid=18573073 |url=http://arjournals.annualreviews.org/doi/abs/10.1146/annurev.biophys.37.032807.125912 |doi=10.1146/annurev.biophys.37.032807.125912}}</ref> In vertebrates, three main groups of actin [[isoforms]], [[ACTA1|alpha]], [[ACTB|beta]], and [[ACTG1|gamma]] have been identified. The alpha actins, found in muscle tissues, are a major constituent of the contractile apparatus. The beta and gamma actins coexist in most cell types as components of the [[cytoskeleton]], and as [[chemical mediator|mediators]] of internal cell motility.

Its [[amino acid sequence]] is also one of the most highly-[[Conserved sequence|conserved]] of the proteins as it has changed little over the course of [[evolution]], differing by no more than 20% in [[species]] as diverse as [[algae]] and [[human]]s. It is therefore considered to have an optimised [[protein structure|structure]]. It has two distinguishing features: it is an [[enzyme]] that slowly [[hydrolysis|hydrolizes]] [[Adenosine triphosphate|ATP]], the "universal energy currency" of biological processes. However, the ATP is required in order to maintain its structural integrity. Its efficient structure is formed by an almost unique [[protein folding|folding]] process. In addition, it is able to carry out more [[protein-protein interaction|interactions]] than any other protein, which allows it to perform a wider variety of functions than other proteins at almost every level of cellular life. [[Myosin]] is an example of a protein that bonds with actin. Another example is [[villin]], which can weave actin into bundles or cut the filaments depending on the concentration of [[calcium]] cations in the surrounding medium.<ref name=biolcel />

A cell’s ability to dynamically form microfilaments provides the scaffolding that allows it to rapidly remodel itself in response to it’s environment or to the organism’s internal [[signal transduction|signals]], for example, to increase cell membrane absorption or increase [[cell adhesion]] in order to form cell [[tissue (biology)|tissue]]. Other enzymes or [[organelle]]s such as [[cillium|cillia]] can be anchored to this scaffolding in order to control the deformation of the external [[cell membrane]], which allows [[endocytosis]] and [[cytokinesis]]. It can also produce movement either by itself or with the help of [[molecular motors]]. Actin therefore contributes to processes such as the intracellular transport of [[Vesicle (biology and chemistry)|vesicles]] and organelles as well as [[muscle|muscular contraction]] and cellular migration. It therefore plays an important role in [[embryogenesis]], the healing of wounds and the invasivity of [[cancer]] cells. The evolutionary origin of actin can be traced to [[prokaryotic cells]], which have equivalent proteins. Lastly, actin plays an important role in the control of [[gene expression]].

A large number of [[disease|illnesses and diseases]] are caused by [[mutation]]s in [[allele]]s of the [[gene]]s that regulate the production of actin or of its associated proteins. The production of actin is also key to the process of [[infection]] by some [[Pathogen|pathogenic]] [[microorganism]]s. Mutations in the different genes that regulate actin production in humans can cause [[Myopathy|muscular diseases]], variations in the size and function of the [[heart]] as well as [[deafness]]. The make-up of the cytoskeleton is also related to the pathogenicity of intracellular [[bacteria]] and [[virus]]es, particularly in the processes related to evading the actions of the [[immune system]].<ref name="alberts">{{cite book| author = Alberts ''et al''| title = Biología molecular de la célula|language = Spanish |year = 2004| publisher = Barcelona: Omega| id = ISBN 54-282-1351-8}}</ref>

== History ==
[[Image:GyorgyiNIH.jpg|thumb|[[Nobel Prize]] winning [[physiologist ]] [[Albert von Szent-Györgyi Nagyrápolt]], co-discoverer of actin with [[Brunó Ferenc Straub]].]]
Actin was first observed [[experiment]]ally in 1887 by [[W.D. Halliburton]], who extracted a protein from muscle that 'coagulated' preparations of myosin that he called "myosin-ferment".<ref name="Halliburton">{{cite journal |author=Halliburton, W.D. |title=On muscle plasma |journal=J. Physiol. |volume=8 |issue= 3-4|pages=133 |year=1887 |pmid=16991477 |pmc=1485127 }}</ref> However, Halliburton was unable to further refine his findings, and the discovery of actin is credited instead to [[Brunó Ferenc Straub]], a young [[Biochemistry|biochemist]] working in [[Albert Szent-Györgyi]]'s laboratory at the Institute of Medical Chemistry at the [[University of Szeged]], [[Hungary]].

In [[1942]], Straub developed a novel technique for [[Extraction (chemistry)|extracting]] muscle protein that allowed him to isolate substantial amounts of relatively [[chemical substance|pure]] actin. Straub's method is essentially the same as that used in [[laboratory|laboratories]] today. Szent-Gyorgyi had previously described the more viscous form of myosin produced by slow muscle extractions as 'activated' myosin, and, since Straub's protein produced the activating effect, it was dubbed ''actin''. Adding [[Adenosine triphosphate|ATP]] to a mixture of both proteins (called ''actomyosin'') causes a decrease in viscosity. The hostilities of [[World War II]] meant Szent-Gyorgyi and Straub were unable to publish the work in [[Western countries|Western]] [[scientific journal]]s. Actin therefore only became well known in the West in 1945, when their paper was published as a supplement to the ''Acta Physiologica Scandinavica''.<ref name="Szent_Gyorgyi">{{cite journal |author=Szent-Gyorgyi, A. |title=Studies on muscle |journal=Acta Physiol Scandinav |volume=9 |issue=Suppl |pages=25 |year=1945 }}</ref> Straub continued to work on actin, and in [[1950]] reported that actin contains bound [[adenosine triphosphate|ATP]] <ref name="Straub">{{cite journal |author=Straub FB, Feuer G |title=Adenosine triphosphate. The functional group of actin. 1950 |journal=Biochim. Biophys. Acta |volume=1000 |issue= |pages=180–95 |year=1989 |pmid=2673365 }}</ref> and that, during [[polymer|polymerization]] of the protein into [[microfilament]]s, the [[nucleotide]] is [[hydrolysis|hydrolyzed]] to [[adenosine diphosphate|ADP]] and inorganic [[phosphate]] (which remain bound to the microfilament). Straub suggested that the transformation of ATP-bound actin to ADP-bound actin played a role in muscular contraction. In fact, this is true only in [[smooth muscle]], and was not supported through experimentation until [[2001]].<ref name="Straub">{{cite journal| author =Straub, F.B. and Feuer, G.| title =Adenosine triphosphate the functional group of actin.| year =1950 | journal = Biochim.Biophys. Acta.| id = PMID 2673365}}</ref><ref name="Bárány">{{cite journal |author=Bárány M, Barron JT, Gu L, Bárány K |title=Exchange of the actin-bound nucleotide in intact arterial smooth muscle |journal=J. Biol. Chem. |volume=276 |issue=51 |pages=48398–403 |year=2001 |month=December |pmid=11602582 |doi=10.1074/jbc.M106227200 |url=http://www.jbc.org/cgi/pmidlookup?view=long&pmid=11602582}}</ref>

The [[peptide sequence|amino acid sequencing]] of actin was completed by M. Elzinga and co-workers in [[1973]],<ref name="Elzinga">{{cite journal | title = Complete amino-acid sequence of actin of rabbit skeletal muscle | url = http://www.pnas.org/cgi/reprint/70/9/2687.pdf | year= 1973 | pub-journal= Proceedings of the National Academy of Sciences | pages = 2687–2691 | volume= 70 | number= 9 | surname1= Elzinga| name1= M. | surname2= Collins| name2= J.H. | surname3= Kuehl| name3= W.M. | surname4= Adelstein | name4= R.S. | access date= 29 June 2009}}</ref>. The [[X-ray crystallography|crystal structure]] of G-actin was solved in 1990 by Kabsch and colleagues.<ref name="Kabsch">{{cite journal |author=Kabsch W, Mannherz HG, Suck D, Pai EF, Holmes KC |title=Atomic structure of the actin:DNase I complex |journal=Nature |volume=347 |issue=6288 |pages=37–44 |year=1990 |month=September |pmid=2395459 |doi=10.1038/347037a0}}</ref> In the same year, a model for F-actin was proposed by Holmes and colleagues following experiments using co-crystallization with different proteins.<ref name="Holmes">{{cite journal |author=Holmes KC, Popp D, Gebhard W, Kabsch W |title=Atomic model of the actin filament |journal=Nature |volume=347 |issue=6288 |pages=44–9 |year=1990 |month=September |pmid=2395461 |doi=10.1038/347044a0}}</ref> The procedure of co-crystallization with different proteins was used repeatedly during the following years, until in [[2001]] the isolated protein was crystallized along with ADP. However, there is still no high-resolution X-ray structure of F-actin. The crystallization of F-actin was possible due to the use of a [[rhodamine]] conjugate that impedes polymerization by blocking the amino acid [[cistein|cys-374]].<ref name="dominguez">{{cite journal| surname =Domínguez|name=R| co-authors =Otterbein LR, Graceffa P| year =2001 | month =July | title =The crystal structure of uncomplexed actin in the ADP state |volume =293 | number =5530 | pages =708-11| pmid =11474115}}</ref> Christine Oriol-Audit died in the same year that actin was first crystalized but she was the researcher that in [[1977]] first crystalized actin in the absence of Actin Binding Proteins (ABPs). However, the resulting crystals were too small for the available technology of the time.<ref name="Oriol">{{cite journal| surname =Oriol | name =C | co-authors=Dubord, C and Landon, F |year=1977 | month =January | title =Crystallization of native striated-muscle actin | journal =FEBS Lett. | volume=73 | number=1 | pages=89-91| pmid=320040 }}</ref>

Although no high-resolution model of actin’s filamentous form currently exists, in [[2008]] Sawaya’s team were able to produce a more exact model of its structure based on multiple crystals of actin [[Dimer (chemistry)|dimers]] that bind in different places.<ref>{{cite journal| author =Sawaya MR, Kudryashov DS, Pashkov I, Adisetiyo H, Reisler E, Yeates TO.| title =Multiple crystal structures of actin dimers and their implications for interactions in the actin filament | year =2008 | journal = Acta Crystallogr D Biol Crystallogr. |id = PMID 18391412}}</ref> This model has subsequently been further refined by Sawaya and Lorenz. Other approaches such as the use of [[cryo-electron microscopy]] and [[synchrotron radiation]] have recently allowed increasing resolution and better understanding of the nature of the interactions and conformational changes implicated in the formation of actin filaments.<ref name="Narita">{{cite journal| surname =Narita | co-authors=Takeda S, Yamashita A, Maéda Y. | year=2006 |month=November | title =Structural basis of actin filament capping at the barbed-end: a cryo-electron microscopy study | journal =Embo J | volume=25 | number=23 | pages=5626-33| doi 10.1038/sj.emboj.7601395| url=http://www.nature.com/emboj/journal/v25/n23/pdf/7601395a.pdf | format =[[PDF]] | access date=[[27 June]] [[2009]]}}</ref><ref name="Toshiro" />

== Structure ==

Actin is one of the most abundant proteins in [[eukaryote]]s, where it is found throughout the cytoplasm.<ref name=biolcel /> In fact, in [[Myocyte|muscle fibres]] it comprises 20% of total cellular protein by weight and between 1% and 5% in other cells. However, there is not only one type of actin, the [[gene]]s that code for actin are defined as a [[gene family]] (a family that in plants contains more than 60 elements, including genes and [[pseudogene]]s and in humans more than 30 elements).<ref name=Ponte1983>{{Cite journal | surname1= Ponte | name1= P. | surname2= Gunning | name2= P. | surname3= Blau | name3= H. | surname4= Kedes | name4= L. | year= 1983 | title = Human actin genes are single copy for alpha-skeletal and alpha-cardiac actin but multicopy for β- and γ-cytoskeletal genes: 3' untranslated regions are isotype specific but are conserved in evolution| pub-journal= Molecular and Cellular Biology | volume= 3 | number= 10 | page = 1783–1791 | url = http://mcb.asm.org/cgi/content/abstract/3/10/1783}}</ref> This means that the genetic information of each individual contains instructions that generate actin variants (called [[isoform]]s) that possess slightly different functions. This, in turn, means that eukaryotic organisms [[gene expresion|express]] different genes that give rise to: α-actin, which is found in contractile structures; β-actin, found at the expanding edge of cells that use the projection of their cellular structures as their means of mobility; and γ-actin, which is found in the filaments of [[stress fibre]]s.<ref name="lodish">{{cite book| author = Lodish et al.| title = Biología celular y molecular| year = 2005| publisher = Buenos Aires: Médica Panamericana | language = Spanish | id = ISBN 950-06-1974-3}}</ref> In addition to the similarities that exist between an organism’s isoforms there is also an [[Evolution|evolutionary conservation]] in the structure and function even between organisms contained in different eukaryotic [[Domain (biology)|domains]]: in [[bacteria]] the actin [[Homology (biology)|homologue]] [[MreB]] has been identified, which is a protein that is capable of polymerizing into microfilaments;<ref name="Toshiro" /> and in [[archaea]] the homologue Ta0583 is even more similar to the eukaryotic actins.<ref>Futoshi Hara, Kan Yamashiro, Naoki Nemoto, Yoshinori Ohta, Shin-ichi Yokobori, Takuo Yasunaga, Shin-ichi Hisanaga, and Akihiko Yamagishi. (2007): [http://jb.asm.org/cgi/content/abstract/189/5/2039 An Actin Homologue of the Archaeon Thermoplasma acidophilum That Retains the Ancient Characteristics of Eukaryotic Actin]. Journal of Bacteriology, p. 2039-2045, Vol. 189, No. 5 doi:10.1128/JB.01454-06</ref>

Cellular actin has two forms: monomeric [[Globular protein|globules]] called G-actin and [[Biopolymer|polymeric]] filaments called F-actin (that is, as filaments made up of many G-actin monomers). F-actin can also be described as a microfilament.
Two parallel F-actin strands must rotate 166 degrees to lay correctly on top of each other. This creates the double helix structure of the microfilaments found in the cytoskeleton. Microfilaments measure approximately 7 [[nanometer|nm]] in diameter with the helix repeating every 37&nbsp;nm.Each strand of actin is bound to a molecule of [[adenosine triphosphate]] (ATP) or [[adenosine diphosphate]] (ADP) that is associated with a [[magnesium|Mg<sup>2+</sup>]] cation. The most commonly found forms of actin, compared to all the possible combinations, are ATP-G-Actin and ADP-F-actin.<ref name=Graceffa2003>{{Cite journal | surname1= Graceffa | name1= Philip | surname2= Dominguez | name2= Roberto | year= 2003 | title = Crystal Structure of Monomeric Actin in the ATP State: STRUCTURAL BASIS OF NUCLEOTIDE-DEPENDENT ACTIN DYNAMICS | pub-journal= Journal of Biological Chemistry | volume= 278 | number= 36 | page = 34172–34180 | doi = 10.1074/jbc.M303689200 | url = http://www.jbc.org/cgi/content/full/278/36/34172 | pmid = 12813032}}</ref><ref name=Reisler1993>{{Cite journal | surname= Reisler | name= E. | year= 1993 | title = Actin molecular structure and function | pub-journal= Curr Opin Cell Biol | volume= 5 | number= 1 | page = 41–7 | doi = 10.1016/S0955-0674(05)80006-7 | url = http://www.ncbi.nlm.nih.gov/pubmed/8448029}}</ref>

=== G-Actin ===

[[Scanning electron microscope]] images indicate that G-actin has a globular structure; however, [[X-ray crystallography]] shows that these globules are comprised of two lobes separated by a cleft. This structure represents the “ATPase fold”, which is a centre of [[Enzyme kinetics|enzymatic catalysis]] that binds ATP and Mg<sup>2+</sup> and hydrolyzes the former to ADP plus [[phosphate]]. This fold is a conserved structural motif that is also found in other proteins that interact with triphosphate [[nucleotide]]s such as [[hexokinase]] (an enzyme used in energy [[metabolism]]) or in [[Hsp70]] proteins (a protein family that play an important part in protein folding).<ref>{{cite web|url = http://www.ncbi.nlm.nih.gov/Structure/cdd/cddsrv.cgi?uid=28896 |title =NCBI Conserved Domains: ATP binding site |access date =[[26 December]] |access year =2008}}</ref> G-actin is only functional when it contains either ADP or ATP in its cleft, however, the form that is bound to ATP predominates in cells when actin is present in its free state.<ref name="Graceffa2003" />
[[Image:Actin-ATP Dominguez.svg|thumb|300px|[[Molecular model|Ribbon model]] of actin extracted from the [[striated muscle tissue]] of a [[European rabbit|rabbit]] after Graceffa and Domínguez, 2003. The four [[Protein secondary structure|subdomains]] can be seen, as well as the [[N-terminus|N]] and [[C-terminus|C]] termini and the position of the ATP bond. The [[molecule]] is orientated using the usual convention of placing the - end (pointed end) in the upper part and the + end (barbed end) in the lower part.<ref name="dominguez" />]]

The [[X-ray crystallography]] model of actin that was produced by Kabsch from the [[striated muscle tissue]] of [[European rabbit|rabbit]]s is the most commonly used in structural studies as it was the first to be [[Separation process|purified]]. The G-actin crystallized by Kabsch is approximately 67 x 40 x 37 [[Angstrom|Å]] in size, has a [[molecular mass]] of 41,785 [[Dalton (unit)|Da]] and an estimated [[isoelectric point]] of 4.8. Its [[Electric charge|net charge]] at [[pH]] = 7 is -7.<ref name="Elzinga2" />
<ref name=autogenerated1>{{cite journal| surname =Elzinga | co-authors =Collins JH, Kuehl WM, Adelstein RS | title= Complete amino-acid sequence of actin of rabbit skeletal muscle.| journal = Proc Natl Acad Sci U S A. | volume=70 | number=9 | url=http://www.pubmedcentral.nih.gov/picrender.fcgi?artid=427084&blobtype=pdf | format = [[PDF]]}}</ref>

;Primary structure

Elzinga and co-workers first determined the complete [[peptide sequence]] for this type of actin in [[1973]], with later work by the same author adding further detail to the model. It contains 374 [[amino acid]] residues. Its [[N-terminus]] is highly [[acid]]ic and starts with an [[Acetyl|acetyled]] [[Aspartic acid|aspartate]] in its amino group. While its [[C-terminus]] is [[Base (chemistry)|alkaline]] and is formed by a [[phenylalanine]] preceded by a [[cysteine]], which has a degree of functional importance. Both extremes are in close proximity within the I-subdomain. An anomalous [[histidine|''N''<sup>τ</sup>-methylhistidine]] is located at position 73.<ref name="Elzinga2">{{cite journal| surname =Elzinga | co-authors =Collins, JH | year=1975 |month=August | title =The primary structure of actin from rabbit skeletal muscle. Completion and analysis of the amino acid sequence| journal =J Biol Chem. | volume=250 | number=15 | url=http://www.ncbi.nlm.nih.gov/pubmed/1150665?dopt=Abstract }}</ref>

;Tertiary structure - domains

The tertiary structure is formed by two [[Protein domain|domains]] known as the large and the small, which are separated by a cleft centred around the location of the bond with [[adenosine triphosfate|ATP]]-[[adenosine diphosphate|ADP]]+[[phosphate|P<sub>i</sub>]]. Below this there is a deeper notch called a “groove”. In the [[Protein#Structure|native state]], despite their names, both have a comparable depth.<ref name="Elzinga" />

The normal convention in [[topology|topological]] studies means that a protein is shown with the biggest domain on the left-hand side and the smallest domain on the right-hand side. In this position the smaller domain is in turn divided into two: subdomain I (lower position, residues 1-32, 70-144 and 338-374) and subdomain II (upper position, residues 33-69). The larger domain is also divided in two: subdomain III (lower, residues 145-180 and 270-337) and subdomain IV (higher, residues 181-269). The exposed areas of subdomains I and III are referred to as the “barbed” ends, while the exposed areas of domains II and IV are termed the “pointed" ends. This nomenclature refers to the fact that, due to the small mass of subdomain II actin is polar; the importance of this will be discussed below in the discussion on assembly dynamics. Some authors call the subdomains Ia, Ib, IIa and IIb, respectively.<ref name="dosremedios" />

;Other important structures

The most notable supersecondary structure is a five chain [[beta sheet]] that is comprised of a β-meander and a β-α-β clockwise unit. It is present in both domains suggesting that the protein arose from gene duplication.<ref name="Kabsch" />
* The [[Adenosine triphosphate|adenosine nucleotide]] binding site is located between two [[beta hairpin]]-shaped structures pertaining to the I and III domains. The residues that are involved are Asp11-Lys18 and Asp154-His161 respectively.
* The [[cation|divalent cation]] binding site is located just below that for the adenosine nucleotide. ''In vivo'' it is most often formed by [[Magnesium|Mg<sup>2+</sup>]] or [[Calcium|Ca<sup>2+</sup>]] while ''in vitro'' it is formed by a chelating structure made up of [[lysine|Lys18]] and two [[oxygen]]s from the nucleotide’s α-and β-[[phosphate]]s. This calcium is coordinated with six water molecules that are retained by the amino acids [[aspartic acid|Asp11]], Asp154, and [[glutamine|Gln137]]. They form a complex with the nucleotide that restricts the movements of the so-called “hinge” region, located between residues 137 and 144. This maintains the native form of the protein until its withdrawal [[Denaturation (biochemistry)|denatures]] the actin monomer. This region is also important because it determines whether the protein’s cleft is in the “open” or "closed” conformation.<ref name="dosremedios" /><ref name="dominguez" />
* It is highly likely that there are at least three other centres with a lesser [[electron affinity|affinity]] (intermediate) and still others with a low affinity for divalent cations. It has been suggested that these centres may play a role in the polymerization of actin by acting during the activation stage.<ref name="dosremedios" />
* There is a structure in subdomain 2 that is called the “D-loop” because it binds with [[DNase I]], it is located between the [[histidine|His40]] and [[glycine|Gly48]] residues. It has the appearance of a disorderly element in the majority of crystals, but it looks like a β-sheet when it is complexed with DNase I. Domínguez “et al.” suggest that the key event in polymerization is probably the propagation of a conformational change from the centre of the bond with the nucleotide to this domain, which changes from a loop to a spiral. However, this theory has been refuted by other studies.<Ref name="dominguez" /><ref name="Rould">{{cite journal| surname =Rould | co-authors= Wan Q, Joel PB, Lowey S, Trybus KM.| year=2006 |month=October | title =Crystal structures of expressed non-polymerizable monomeric actin in the ADP and ATP states | journal =J Biol chem | volume=281 | number=42 | doi=10.1074/jbc.M601973200| url=http://www.jbc.org/cgi/content/full/281/42/31909?view=long&pmid=16920713}}</ref>

=== F-Actin ===
The classical description of F-actin states that is has a filamentous structure that can be considered to be a single stranded [[Fleming’s left-hand rule for motors|levorotatory]] [[helix]] with a rotation of 166º around the helical axis and an axial translation of 27.5 [[angstrom|Å]], or a single stranded [[Right-hand rule|dextrorotatory]] helix with a cross over spacing of 350-380 Å, with each actin surrounded by four more.<ref name=Devlin>{{cite book |surnames=Devlin |name=Thomas M |title=Bioquímica: Libro de texto con aplicaciones clínicas |url=http://books.google.com/books?id=p3DCb9lTLx8C&pg=PA1021&lpg=PA1021&dq=la+miosina+se+une+a+la+actina+residuo&source=bl&ots=YZZKRVRbRQ&sig=WqMv1TZlXS6hakU88tdHgNLfbt4&hl=es&ei=9yZHSrjVBuPTjAeF4aRk&sa=X&oi=book_result&ct=result&resnum=4 |edition=4|year=2004|publisher=Reverte |language=Spanish |isbn=8429172084|pages=1021 |chapter=23}}.</ref> The symmetry of the actin polymer at 2.17 subunits per turn of a helix is incompatible with the formation of [[crystal]]s, which is only possible with a symmetry of exactly 2, 3, 4 or 6 subunits per turn. Therefore, [[Molecular modelling|models]] have to be constructed that explain these anomalies using data from [[Electron microscope|electron microscopy]], [[cryo-electron microscopy]], crystallization of dimers in different positions and [[X-ray cristalography|diffraction of X-rays]].<ref name="Toshiro">Toshiro Oda, Mitsusada Iwasa, Tomoki Aihara, Yuichiro Maéda and Akihiro Narita (2009): [http://www.nature.com/nature/journal/v457/n7228/abs/nature07685.html The nature of the globular- to fibrous-actin transition]. Nature 457, 441-445 (22 January 2009) | doi:10.1038/nature07685</ref> It should be pointed out that it is not correct to talk of a “structure” for a molecule as dynamic as the actin filament. In reality we talk of distinct structural states, in these the measurement of axial translation remains constant at 27.5 Å while the subunit rotation data shows considerable variability, with displacements of up to 10% from its optimum position commonly seen. Some proteins, such as [[cofilin]] appear to increase the angle of turn, but again this could be interpreted as the establishment of different "structural states". These could be important in the polymerization process.<ref name="Egelman" />

There is less agreement regarding measurements of the turn radius and filament thickness: while the first models assigned a longitude of 25 Å, current X-ray diffraction data, backed up by cryo-electron microscopy suggests a longitude of 23.7 Å. These studies have shown the precise contact points between monomers. Some are formed with units of the same chain, between the "barbed" end on one monomer and the "pointed" end of the next one. While the monomers in adjacent chains make lateral contact through projections from subdomain IV, with the most important projections being those formed by the C-terminus and the hydrophobic link formed by three bodies involving residues 39-42, 201-203 and 286. This model suggests that a filament is formed by monomers in a "sheet" formation, in which the subdomains turn about themselves, this form is also found in the bacterial actin homologue [[MreB]].<ref name="Toshiro" />

The F-actin polymer is considered to have structural polarity due to the fact that all the microfilament’s subunits point towards the same end. This gives rise to a naming convention: the end that possesses an actin subunit that has it’s ATP binding site exposed is called the «(-) end», while the opposite end where the cleft is directed at a different adjacent monomer is called the «(+) end».<ref name="lodish" /> The terms «pointed» and «barbed» referring to the two ends of the microfilaments derive from their appearance under [[transmission electron microscopy]] when samples are examined following a preparation technique called «decoration». This method consists of the addition of [[myosin]] S1 fragments to tissue that has been fixed with [[tannic acid]]. This myosin forms polar bonds with actin monomers, giving rise to a configuration that looks like arrows with feather fletchings along its shaft, where the shaft is the actin and the fletchings are the myosin. Following this logic, the end of the microfilament that does not have any protruding myosin is called the point of the arrow (- end) and the other end is called the barbed end (+ end).<ref>DA Begg, R Rodewald and LI Rebhun (1978): [http://jcb.rupress.org/cgi/content/abstract/79/3/846 The visualization of actin filament polarity in thin sections. Evidence for the uniform polarity of membrane-associated filaments]. The Journal of Cell Biology, Vol 79, 846-852.</ref>
A S1 fragment is composed of the head and neck domains of myosin II. Under physiological conditions, G-actin (the [[monomer]] form) is transformed to F-actin (the [[polymer]] form) by ATP, where the role of ATP is essential.<ref>Histologi by Finn Geneser p. 105. Published by Munksgaard, 1981 [http://books.google.es/books/about/Histologi.html?id=-C3MOAAACAAJ&redir_esc=y] </ref>


The helical F-actin filament found in muscles also contains a [[tropomyosin]] molecule, which is a 40 [[nanometre]] long protein that is wrapped around the F-actin helix. During the resting phase the tropomyosin covers the actin’s active sites so that the actin-myosin interaction cannot take place and produce muscular contraction. There are other protein molecules bound to the tropomyosin thread, these are the [[troponin]]s that have three polymers: [[troponin I]], [[troponin T]] and [[troponin C]].<ref name=fisiologia>Arthur C. Guyton, John E. Hall [http://books.google.es/books?id=K8-d-KzxvTYC Tratado de fisiología médica] (in Spanish). Published by Elsevier España, 2007; page 76. ISBN 84-8174-926-5</ref>

=== Folding ===
[[Image:Prefoldin.png|thumb|[[Molecular model|Ribbon model]] obtained using the [[PyMOL]] programme on [[X-ray cristalography|cristalographs]] of the [[prefoldin]] proteins found in the [[Archaea|arquea]] microorganism ''[[Pyrococcus horikoshii]]''. The six supersecondary structures are present in a coiled helix “hanging” from the central [[beta barrel]]s. These are often compared in the literature to the [[tentacle]]s of a [[jellyfish]]. As far as is visible using [[Electron microscope|electron microscopy]], [[Eukaryote|eukariotic]] prefoldin has a similar structure.<ref name="Simons" />]]

Actin can spontaneously acquire a large part of its [[Protein tertiary structure|tertiary structure]].<ref name="Martinbenito">{{cite journal| surname =Martín-Benito | co-authors =Boskovic J, Gómez-Puertas P, Carrascosa JL, Simons CT, Lewis SA, Bartolini F, Cowan NJ, Valpuesta JM. | year=2002 |month=December | title =Structure of eukaryotic prefoldin and of its complexes with unfolded actin and the cytosolic chaperonin CCT | journal =EMBO J | volume=21 | number=23 | pages=6377-86| pmid=12456645| doi=10.1093/emboj/cdf640 |url=http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pubmed&pubmedid=12456645 }}</ref> However, the way it acquires its [[Protein folding|fully functional form]] from its newly [[Translation (biology)|synthesized]] native form is special and almost unique in protein chemistry. The reason for this special route could be the need to avoid the presence of incorrectly folded actin monomers, which could be toxic as they can act as inefficient polymerization terminators. Nevertheless, it is key to establishing the stability of the cytoskeleton, and additionally, it is an essential process for coordinating the [[cell cycle]].<ref name="Vandamme" /><ref name="Brackley" />

CCT is required in order to ensure that folding takes place correctly. CCT is a group II [[cytosol|cytosolic]] [[Chaperone (protein)|molecular chaperone]] (or chaperonin, a protein that assists in the folding of other macromolecular structures). CCT is formed of a double ring of eight different subunits (hetero-octameric) and it differs from other molecular chaperones, particularly from its homologue [[GroEL]] which is found in [[Archaea|arqueas]], as it does not require a co-chaperone to act as a lid over the central [[Catalysis|catalytic]] cavity. Substrates bind to CCT through specific domains. It was initially thought that it only bound with actin and [[tubulin]], although recent [[immunoprecipitation]] studies have shown that it interacts with a large number of [[polypeptide]]s, which possibly function as [[Enzyme substrate (biology)|substrates]]. It acts through ATP-dependent conformational changes that on occasion require several rounds of liberation and catalysis in order to complete a reaction.<ref name="Stirling">{{cite journal| surname =Stirling| name =PC| co-authors= Cuéllar J, Alfaro GA, El Khadali F, Beh CT, Valpuesta JM, Melki R, Leroux MR | year=2006 |month=March | title =PhLP3 modulates CCT-mediated actin and tubulin folding via ternary complexes with substrates | journal =J Biol Chem | volume=281 | number=11 | pages=7012-21| pmid=16415341 |doi=10.1074/jbc.M513235200 | url=http://www.jbc.org/cgi/content/full/281/11/7012?view=long&pmid=16415341 }}</ref>

In order to successfully complete their folding, both actin and tubulin need to interact with another protein called [[prefoldin]], which is a heterohexameric complex (formed by six distinct subunits), in an interaction that is so specific that the molecules have [[Coevolution|coevolved]]. Actin complexes with prefoldin while it is still being formed, when it is approximately 145 [[amino acid]]s long, specifically those at the N-terminal.<ref name="Hansen">{{cite journal| surname =Hansen | name =WJ| co-authors=Cowan NJ, Welch WJ.| year=1999 |month=April | title =Prefoldin-nascent chain complexes in the folding of cytoskeletal proteins | journal =J Cell Biol. | number=145 | pages=2 | publisher =265-7| pmid=10209023 | url=http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pubmed&pubmedid=10209023 }}</ref>

Different recognition sub-units are used for actin or tubulin although there is some overlap. In actin the subunits that bind with prefoldin are probably PFD3 and PFD4, which bind in two places one between residues 60-79 and the other between residues 170-198. The actin is recognized, loaded and delivered to the cytosolic chaperonin (CCT) in an open conformation by the inner end of prefoldin’s "tentacles” (see the image and note).<ref group=n.>The following link [http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=136944&rendertype=figure&id=cdf640f2] shows a model of prefoldin with actin embedded between the "tentacles of its subunits.</ref> The contact when actin is delivered is so brief that a tertiary complex is not formed, immediately freeing the prefoldin.<ref name="Simons">{{cite journal| surname =Simons | name =CT| co-authors= Staes A, Rommelaere H, Ampe C, Lewis SA, Cowan NJ | year=2004 |month=February | title =Selective contribution of eukaryotic prefoldin subunits to actin and tubulin binding | journal =J Biol Chem. | volume=279 | number=6 | pages=4196-203| pmid=14634002 | doi=10.1074/jbc.M306053200 | url=http://www.jbc.org/cgi/content/full/279/6/4196?view=long&pmid=14634002#REF21 }}</ref>

[[Image:CCT gamma apical.png|thumb|left|Ribbon model of the apical γ-domain of the [[Chaperone (protein)|chaperonin]] CCT.]]

The CCT then causes actin's sequential folding by forming bonds with its subunits rather than simply enclosing it in its cavity.<ref group=n.>The following link from the work of Jaime Martín-Benito and José María Valpuesta from the Spanish CSIC’s National Biotechnology Centre shows the double ring configuration of the CCT chaperonin along with its subunits: [http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=1808031&rendertype=figure&id=f4].</ref> This is why it possesses specific recognition areas in its apical β-domain. The first stage in the folding consists of the recognition of residues 245-249. Next, other determinants establish contact.<ref name="Neirynck">{{cite journal| surname =Neirynck | co-authors= Waterschoot D, Vandekerckhove J, Ampe C, Rommelaere H | year=2006 |month=January | title =Actin interacts with CCT via discrete binding sites: a binding transition-release model for CCT-mediated actin folding | journal =J Mol Biol. | volume=355 | number=1 | pages=124-38 | pmid=16300788}}</ref> Both actin and tubulin bind to CCT in open conformations in the absence of ATP. In actin’s case, two subunits are bound during each conformational change, whereas for tubulin binding takes place with four subunits. Actin has specific binding sequences, which interact with the δ and β-CCT subunits or with δ-CCT and ε-CCT. After AMP-PNP is bound to CCT the substrates move within the chaperonin’s cavity. It also seems that in the case of actin, the [[CAP (protein)|CAP protein]] is required as a possible cofactor in actin's final folding states.<ref name="Brackley">{{cite journal| surname =Brackley | co-authors=Grantham J| year=2009 |month=January | title =Activities of the chaperonin containing TCP-1 (CCT): implications for cell cycle progression and cytoskeletal organisation | journal =Cell Stress Chaperones | volume= 14| number=1 | pages=23-31| pmid=18595008| doi=10.1007/s12192-008-0057-x | url=http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pubmed&pubmedid=18595008 }}</ref>

The exact manner by which this process is regulated is still not fully understood, but it is known that the protein PhLP3 (a protein similar to [[phosducin]]) inhibits its activity through the formation of a tertiary complex.<ref name="Stirling" />

=== ATPase’s catalytic mechanism ===

Actin is an [[ATPase]], which means that it is an [[enzyme]] that [[hydrolysis|hydrolyzes]] ATP. This group of enzymes is characterised by their slow reaction rates. It is know that this ATPase is “active”, that is, its speed increases by some 40,000 times when the actin forms part of a filament.<ref name="Egelman" /> A reference value for this rate of hydrolysis under ideal conditions is around 0.3 [[second|s<sup>-1</sup>]]. Then, the P<sub>i</sub> remains bound to the actin next to the ADP for a long time, until it is liberated next to the end of the filament.<ref name="Vavylonis">{{cite journal| surname =Vavylonis | name =D | co-authors=Yang Q, O'Shaughnessy B. | year=2005 |month=June | title =Actin polymerization kinetics, cap structure, and fluctuations | journal =Proc Natl Acad Sci U S A. | volume=102 | number=24 | pages=8543-8| pmid=15939882| doi=10.1073/pnas.0501435102| url=http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pubmed&pubmedid=15939882 }}</ref>

The exact molecular details of the catalytic mechanism are still not fully understood. Although there is much debate on this issue, it seems certain that a "closed" conformation is required for the hydrolysis of ATP, and it is thought that the residues that are involved in the process move to the appropriate distance.<ref name="Egelman">{{cite journal| surname =Egelman | co-authors=Reisler, E | year=2007 |month=December | title =Actin structure and function: what we still do not understand | journal =J Biol Chem | volume=282 | number=50 | doi=10.1074/jbc.R700030200 | url=http://www.jbc.org/cgi/content/full/282/50/36133 }}</ref> The [[glutamic acid]] Glu137 is one of the key residues, which is located in subdomain 1. Its function is to bind the water molecule that produces a [[nucleophile|nucleophilic attack]] on the ATP’s γ-phosphate [[chemical bond|bond]], while the nucleotide is strongly bound to subdomains 3 and 4. The slowness of the catalytic process is due to the large distance and skewed position of the water molecule in relation to the reactant. It is highly likely that the conformational change produced by the rotation of the domains between actin’s G and F forms moves the Glu137 closer allowing its hydrolysis. This model suggests that the polymerization and ATPase’s function would be decoupled straight away.<ref name="Toshiro" />

==Genetics==
[[Image:Adherens Junctions structural proteins.svg|thumb|350px|]]Principal interactions of structural proteins are at [[cadherin]]-based adherens junction. Actin filaments are linked to α-[[actinin]] and to the membrane through [[vinculin]]. The head domain of vinculin associates to E-cadherin via α-, β-, and γ-catenins. The tail domain of vinculin binds to membrane lipids and to actin filaments.
Actin has been one of the most highly conserved proteins throughout evolution because it interacts with a large number of other proteins. It has 80.2% sequence [[Conservation (genetics)|conservation]] at the [[gene]] level between ''[[Human|Homo sapiens]]'' and ''[[Saccharomyces cerevisiae]]'' (a species of yeast), and 95% conservation of the [[primary structure]] of the protein product.


Although most [[yeast]]s have only a single actin gene, higher [[eukaryote]]s, in general, [[gene expression|express]] several [[isoform]]s of actin encoded by a family of related genes. [[Mammal]]s have at least six actin isoforms coded by separate genes,<ref name="Vandekerckhove J. and Weber K.">{{cite journal |doi=10.1016/0022-2836(78)90020-7 |author=Vandekerckhove J, Weber K |title=At least six different actins are expressed in a higher mammal: an analysis based on the amino acid sequence of the amino-terminal tryptic peptide |journal=J. Mol. Biol. |volume=126 |issue=4 |pages=783–802 |year=1978 |month=December |pmid=745245 |url=http://linkinghub.elsevier.com/retrieve/pii/0022-2836(78)90020-7}}</ref> which are divided into three classes (alpha, [[Beta-actin|beta]] and gamma) according to their [[isoelectric point]]s. In general, alpha actins are found in muscle (α-skeletal, α-aortic smooth, α-cardiac, and γ2-enteric smooth), whereas beta and gamma isoforms are prominent in nonmuscle cells (β- and γ1-cytoplasmic). Although the amino acid sequences and ''[[in vitro]]'' properties of the isoforms are highly similar, these isoforms cannot completely substitute for one another ''[[in vivo]]''.<ref name="Khaitlina SY">{{cite journal |doi=10.1016/S0074-7696(01)02003-4 |author=Khaitlina SY |title=Functional specificity of actin isoforms |journal=Int. Rev. Cytol. |volume=202 |issue= |pages=35–98 |year=2001 |pmid=11061563 }}</ref>
Although most [[yeast]]s have only a single actin gene, higher [[eukaryote]]s, in general, [[gene expression|express]] several [[isoform]]s of actin encoded by a family of related genes. [[Mammal]]s have at least six actin isoforms coded by separate genes,<ref name="Vandekerckhove J. and Weber K.">{{cite journal |doi=10.1016/0022-2836(78)90020-7 |author=Vandekerckhove J, Weber K |title=At least six different actins are expressed in a higher mammal: an analysis based on the amino acid sequence of the amino-terminal tryptic peptide |journal=J. Mol. Biol. |volume=126 |issue=4 |pages=783–802 |year=1978 |month=December |pmid=745245 |url=http://linkinghub.elsevier.com/retrieve/pii/0022-2836(78)90020-7}}</ref> which are divided into three classes (alpha, [[Beta-actin|beta]] and gamma) according to their [[isoelectric point]]s. In general, alpha actins are found in muscle (α-skeletal, α-aortic smooth, α-cardiac, and γ2-enteric smooth), whereas beta and gamma isoforms are prominent in non-muscle cells (β- and γ1-cytoplasmic). Although the amino acid sequences and ''[[in vitro]]'' properties of the isoforms are highly similar, these isoforms cannot completely substitute for one another ''[[in vivo]]''.<ref name="Khaitlina SY">{{cite journal |doi=10.1016/S0074-7696(01)02003-4 |author=Khaitlina SY |title=Functional specificity of actin isoforms |journal=Int. Rev. Cytol. |volume=202 |issue= |pages=35–98 |year=2001 |pmid=11061563 }}</ref>


The typical actin gene has an approximately 100-nucleotide [[5' UTR]], a 1200-nucleotide [[Translation (genetics)|translated]] region, and a 200-nucleotide [[3' UTR]]. The majority of actin genes are interrupted by [[intron]]s, with up to six introns in any of 19 well-characterised locations. The high conservation of the family makes actin the favoured model for studies comparing the introns-early and introns-late models of intron evolution.
The typical actin gene has an approximately 100-nucleotide [[5' UTR]], a 1200-nucleotide [[Translation (genetics)|translated]] region, and a 200-nucleotide [[3' UTR]]. The majority of actin genes are interrupted by [[intron]]s, with up to six introns in any of 19 well-characterised locations. The high conservation of the family makes actin the favoured model for studies comparing the introns-early and introns-late models of intron evolution.


All nonspherical [[prokaryote]]s appear to possess genes such as [[MreB]], which encode [[homology (biology)|homologues]] of actin; these genes are required for the cell's shape to be maintained. The [[plasmid]]-derived gene ParM encodes an actin-like protein whose polymerised form is [[Microtubule#Dynamic instability|dynamically unstable]], and appears to partition the plasmid [[DNA]] into the daughter cells during cell division by a mechanism analogous to that employed by microtubules in eukaryotic [[mitosis]].<ref name=" Garner EC et al.">{{cite journal |author=Garner EC, Campbell CS, Weibel DB, Mullins RD |title=Reconstitution of DNA segregation driven by assembly of a prokaryotic actin homolog |journal=Science |volume=315 |issue=5816 |pages=1270–4 |year=2007 |month=March |pmid=17332412 |pmc=2851738 |doi=10.1126/science.1138527 |url=http://www.sciencemag.org/cgi/pmidlookup?view=short&pmid=17332412}}</ref>
All non-spherical [[prokaryote]]s appear to possess genes such as [[MreB]], which encode [[homology (biology)|homologues]] of actin; these genes are required for the cell's shape to be maintained. The [[plasmid]]-derived gene ParM encodes an actin-like protein whose polymerized form is [[Microtubule#Dynamic instability|dynamically unstable]], and appears to partition the plasmid [[DNA]] into its daughter cells during cell division by a mechanism analogous to that employed by microtubules in eukaryotic [[mitosis]].<ref name=" Garner EC et al.">{{cite journal |author=Garner EC, Campbell CS, Weibel DB, Mullins RD |title=Reconstitution of DNA segregation driven by assembly of a prokaryotic actin homolog |journal=Science |volume=315 |issue=5816 |pages=1270–4 |year=2007 |month=March |pmid=17332412 |pmc=2851738 |doi=10.1126/science.1138527 |url=http://www.sciencemag.org/cgi/pmidlookup?view=short&pmid=17332412}}</ref>
Actin is found in both smooth and rough endoplasmic reticulums.
Actin is found in both smooth and rough endoplasmic reticulums.


==Functions==
Actin forms [[microfilament]]s which are typically one of the most dynamic of the three subclasses of the eukaryotic [[cytoskeleton]].


== Assembly dynamics ==
In turn, this gives actin major functions in cells:
[[Image:Thin filament formation.svg|thumb|centre|Thin filament formation showing the polymerization mechanism for converting G-actin to F-actin; note the hydrolysis of the ATP.]]
* To form [[microfilament]]s to give mechanical support to cells, and provide trafficking routes through the cytoplasm to support signal transduction
===Nucleation and polymerization===
* To allow [[cell motility]] in cells which undergo [[amoeboid]] motion using [[pseudopods]] (see [[Microfilament#Microfilament-based motility by actoclampin molecular motors|actoclampin molecular motors]]) and [[phagocytosis]], for example of bacteria by [[macrophage]]s

Actin polymerization and depolymerization is necessary in [[chemotaxis]] and [[cytokinesis]]. Nucleating factors are necessary to stimulate actin polymerization. One such nucleating factor is the Arp2/3 complex, which mimics a G-actin dimer in order to stimulate the nucleation of G-actin (or monomeric actin). The [[Arp2/3 complex]] binds to actin filaments at 70 degrees to form new actin branches off existing actin filaments. Also, actin filaments themselves bind ATP, and hydrolysis of this ATP stimulates destabilization of the polymer.

The growth of actin filaments can be regulated by [[thymosin]] and [[profilin]]. Thymosin binds to G-actin to buffer the polymerizing process, while profilin binds to G-actin to exchange [[Adenosine diphosphate|ADP]] for [[Adenosine triphosphate|ATP]], promoting the monomeric addition to the barbed, plus end of F-actin filaments.

F-actin is both [[Strength of materials|strong]] and dynamic. Unlike other [[polymer]]s, such as [[DNA]], whose constituent elements are bound together with [[covalent bond]]s, the monomers of actin filaments are assembled by weaker bonds. The lateral bonds with neighbouring monomers resolve this anomaly, which in theory should weaken the structure as they can be broken by thermal agitation. In addition, the weak bonds give the advantage that the filament ends can easily release or incorporate monomers. This means that the filaments can be rapidly remodelled and can change cellular structure in response to an environmental stimulus. Which, along with the [[biochemical]] mechanism by which it is brought about is known as the "assembly dynamic".<ref name="alberts" />

; ''In vitro'' studies

Studies focussing on the accumulation and loss of subunits by microfilaments are carried out ''[[in vitro]]'' (that is, in the laboratory and not on cellular systems) as the polymerization of the resulting actin gives rise to the same F-actin as produced ''[[in vivo]]''. The ''in vivo'' process is controlled by a multitude of proteins in order to make it responsive to cellular demands, this makes it difficult to observe its basic conditions.<ref name=Kawamura1970>{{Cite journal | surname1= Kawamura | name1= M. | surname2= Maruyama | name2= K. | year= 1970 | title = Electron Microscopic Particle Length of F-Actin Polymerized in Vitro | pub-journal= Journal of Biochemistry | volume= 67 | number= 3 | page = 437 | url = http://jb.oxfordjournals.org/cgi/content/abstract/67/3/437}}</ref> ''In vitro'' production takes place in a sequential manner: first, there is the «activation phase», when the bonding and exchange of divalent cations occurs in specific places on the G-actin, which is bound to ATP, this produces a conformational change, sometimes called G*-actin or F-actin monomer as it is very similar to the units that are located on the filament.<ref name="dosremedios" /> This prepares it for the «nucleation phase», in which the G-actin gives rise to small unstable fragments of F-actin that are able to polymerize. Unstable dimers and trimers are initially formed. The «elongation phase» begins when there are a sufficiently large number of these short polymers. In this phase the filament forms and rapidly grows through the reversible addition of new monomers at both extremes.<ref name=cellcooper>{{cite book|surnames= Cooper|name= Geoffrey M.
|co-authors=Robert E. Hausman|title= The cell: a molecular approach||year= 2007
|publisher= ASM Press, Washington|isbn= 0878932194|chapter=Chapter 12: The Cytoskeleton and Cell Movement}}</ref> Finally, a «stationary equilibrium» is achieved where the G-actin monomers are exchanged at both ends of the microfilament without any change to its total length.<ref name=biolcel>Marc Maillet [http://books.google.es/books?id=54vSCCv33pYC Biología celular] (in Spanish). Published by Elsevier España, 2002; page 132. ISBN 84-458-1105-3</ref> In this last phase the «critical concentration C<sub>c</sub>» is defined as the ratio between the assembly constant and the [[dissociation constant]] for G-actin, where the dynamic for the addition and elimination of dimers and trimers does not produce a change in the microfilament's length. Under normal “in vitro” conditions C<sub>c</sub> is 0.1 μM,<ref>{{cite web| url= http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=1304143&rendertype=table&id=tbl2 |title= Table of the [[equilibrium constants|constants]] of association and dissociation for the different types of monomers to the actin filament, according to the scientific literature.}}</ref> which means that at higher values polymerization occurs and at lower values depolymerization occurs.<ref name=Kirschner1980>{{Cite journal | surname= Kirschner | name= M.W. | year= 1980 | title = Implications of treadmilling for the stability and polarity of actin and tubulin polymers in vivo | pub-journal= The Journal of Cell Biology | volume= 86 | number= 1 | page = 330–334 |10.1083/jcb.86.1.330 | url = http://www.jcb.org/cgi/reprint/86/1/330.pdf | pmid = 6893454 | format= w}}</ref>

;Role of ATP hydrolysis

As indicated above, although actin hydrolyzes ATP, everything points to the fact that ATP is not required for actin to be assembled, given that, on one hand, the hydrolysis mainly takes place inside the filament, and on the other hand the ADP could also instigate polymerization. This poses the question of understanding which [[thermodynamics|thermodynamically]] unfavourable process requires such a prodigious expenditure of [[energy]]. The so-called “actin cycle”, which couples ATP hydrolysis to actin polymerization, consists of the preferential addition of G-actin-ATP monomers to a filament’s barbed end, and the simultaneous disassembly of F-actin-ADP monomers at the pointed end where the ADP is subsequently changed into ATP, thereby closing the cycle, this aspect of actin filament formation is known as “treadmilling”.

ATP is hydrolysed relatively rapidly just after the addition of a G-actin monomer to the filament. There are two hypotheses regarding how this occurs; the [[stochastic]], which suggests that hydrolysis randomly occurs in a manner that is in some way influenced by the neighbouring molecules; and the vectoral, which suggests that hydrolysis only occurs adjacent to other molecules whose ATP has already been hydrolysed. In either case, the resulting P<sub>i</sub> is not released, it remains for some time [[noncovalent bonding|noncovalently]] bound to actin’s ADP, in this way there are three species of actin in a filament: ATP-Actin, ADP+P<sub>i</sub>-Actin and ADP-Actin.<ref name="Vavylonis" /> The amount of each one of these species present in a filament depends on its length and state: as elongation commences the filament has an approximately equal amount of actin monomers bound with ATP and ADP+P<sub>i</sub> and a small amount of ADP-Actin at the (-) end. As the stationary state is reached the situation reverses, with ADP present along the majority of the filament and only the area nearest the (+) end containing ADP+P<sub>i</sub> and with ATP only present at the tip.<ref name=”Lewin”>{{cite book |surnames=Lewin |name=Benjamin |title=Cells |url=http://books.google.com/books?id=2VEGC8j9g9wC&pg=PA378&dq=hydrolysis+actin+polymerization&lr=&client=iceweasel-a&hl=es |format=[[Google]] books|year=2006 |publisher=Jones & Bartlett Publishers |isbn=9780763739058}}</ref>
If we compare the filaments that only contain ADP-Actin with those that include ATP, in the former the critical constants are similar at both ends, while C<sub>c</sub> for the other two nucleotides is different: At the (+) end Cc<sup>+</sup>=0.1 μM, while at the (-) end Cc<sup>-</sup>=0.8 μM, which gives rise to the following situations:<ref name="lodish" />

* For G-actin-ATP concentrations less than Cc<sup>+</sup> no elongation of the filament occurs.
* For G-actin-ATP concentrations less than Cc<sup>-</sup> but greater than Cc<sup>+</sup> elongation occurs at the (+) end.
* For G-actin-ATP concentrations greater than Cc<sup>-</sup> the microfilament grows at both ends.

It is therefore possible to deduce that the energy produced by hydrolysis is used to create a true “stationary state”, that is a flux, instead of a simple equilibrium, one that is dynamic, polar and attached to the filament. This justifies the expenditure of energy as it promotes essential biological functions.<ref name="Vavylonis" /> In addition, the configuration of the different monomer types is detected by actin binding proteins, which also control this dynamism, as will be described in the following section.

Microfilament formation by treadmilling has been found to be atypical in [[stereocilia]]. In this case the control of the structure's size is totally apical and it is controlled in some way by gene expression, that is, by the total quantity of protein monomer synthesized in any given moment.<ref>{{cite journal | doi = 10.1038/nature10745 | issn = 0028-0836, 1476-4687 | volume = 481 | number = 7382 | pages = 520-524 | surname = Zhang | name = Duan-Sun | co-authors = Valeria Piazza, Benjamin J. Perrin, Agnieszka K. Rzadzinska, J. Collin Poczatek, Mei Wang, Haydn M. Prosser, James M. Ervasti, David P. Corey, Claude P. Lechene | title = Multi-isotope imaging mass spectrometry reveals slow protein turnover in hair-cell stereocilia | journal = Nature | access date = 27-01-2012 | date = 15-01-2012 | url = http://www.nature.com/doifinder/10.1038/nature10745 }}</ref>

=== Associated proteins ===

[[Image:Profilin actin complex.png|thumb|An actin (green) - profilin (blue) complex.<ref>[http://www.pdb.org/pdb/explore/explore.do?structureId=2BTF THE STRUCTURE OF CRYSTALLINE PROFILIN-BETA-ACTIN] [[Protein Data Bank]]</ref> The profilin shown belongs to group II, normally present in the [[kidney]]s and the [[brain]].]]

The actin cytoskeleton ''[[in vivo]]'' is not exclusively composed of actin, other proteins are required for its formation, continuance and function. These proteins are called ''[[actin-binding proteins']]'' (ABP) and they are involved in actin’s polymerization, depolymerization, stability, organisation in bundles or networks, fragmentation and destruction.<ref name=biolcel /> The diversity of these proteins is such that actin is thought to be the protein that takes part in the greatest number of [[protein-protein interaction]]s.<ref name="">{{cite journal| surname =Domínguez | name =R| year=2004 |month=November | title =Actin-binding proteins--a unifying hypothesis. | Trends Biochem Sci | volume=29 | number=11 | pages=572-8 |pmid=15501675}}</ref> For example, G-actin sequestering elements exist that impede its incorporation into microfilaments. There are also proteins that stimulate its polymerization or that give complexity to the synthesizing networks.<ref name="lodish" />
* [[Thymosins|Thymosin β-4]] is a 5 kDa protein that can bind with G-actin-ATP in a 1:1 [[stoichiometry]]; which means that one unit of thymosin β-4 binds to one unit of G-actin. Its role is to impede the incorporation of the monomers into the growing polymer.<ref name=Goldschmidt-clermont1992>{{Cite journal | surname1= Goldschmidt-clermont | name1= P.J. | surname2= Furman | name2= M.I. | surname3= Wachsstock | name3= D. | surname4= Safer | name4= D. | surname5= Nachmias | name5= V.T. | surname6= Pollard | name6= T.D. | year= 1992 | title = The control of actin nucleotide exchange by thymosin β-4 and profilin. A potential regulatory mechanism for actin polymerization in cells | pub-journal= Molecular Biology of the Cell | volume= 3 | number= 9 | page = 1015–1024 | url = http://www.molbiolcell.org/cgi/content/abstract/3/9/1015}}</ref>
* [[Profilin]], is a [[cytosol|cytosolic]] protein with a molecular weight of 15 kDa, which also binds with G-actin-ATP with a stoichiometry of 1:1, but it has a different function as it facilitates the replacement of ATP nucleotides by ADP. It is also implicated in other cellular functions, such as the binding of [[prolin]] repetitions in other proteins or of lipids that act as [[second messenger system|secondary messengers]].<ref name="Witke">Witke, W., Podtelejnikov, A., Di Nardo, A., Sutherland, J., Gurniak, C., Dotti, C., and M. Mann (1998) In Mouse Brain Profilin I and Profilin II Associate With Regulators of the Endocytic Pathway and Actin Assembly. The EMBO Journal 17(4): 967-976 {{Entrez Pubmed|9463375}}</ref><ref name="Carlsson">Carlsson L, Nyström LE, Sundkvist I, Markey F, Lindberg U. (1977) Actin polymerizability is influenced by profilin, a low molecular weight protein in non-muscle cells. J. Mol. Biol. 115:465-483 {{Entrez Pubmed|563468}}</ref>

[[Image:Gelsolin.png|thumb|left|The protein [[gelsolin]], which is a key regulator in the assembly and disassembly of actin. It has six subdomains, S1-S6, each of which is composed of a five-stranded [[β-sheet]] flanked by two [[α-helix|α-helices]], one positioned perpendicular to the strands and the other in a parallel position. Both the N-terminal end, (S1-S3), and the C-terminal end, (S4-S6), form an extended β-sheet.<ref name="Kiselar">{{cite journal |author=Kiselar, J., Janmey, P., Almo, S., Chance, M. |title=Visualizing the Ca2+-dependent activation of gelsolin by using synchrotron footprinting |journal=PNAS |year=2003 |volume=100 |number=7 |pages=3942–3947 |url=http://www.pnas.org/cgi/content/full/100/7/3942 |pmid=12655044 |doi=10.1073/pnas.0736004100}}</ref>]]

Other proteins that bind to actin regulate the length of the microfilaments by cutting them, which gives rise to new active ends for polymerization. So that, if a microfilament, that has two ends that monomers can be added to or taken away from, is cut twice, there will be three new microfilaments with six ends. This new situation will favour the dynamics of assembly and disassembly. The most notable of these proteins are [[gelsolin]] and [[cofilin]]. These proteins first achieve a cut by binding to an actin monomer located in the polymer they then change the actin monomer’s [[conformation]] while remaining bound to the newly generated (+) end. This has the effect of impeding the addition or exchange of new G-actin subunits. Depolymerization is encouraged as the (-) ends are not linked to any other molecule.<ref name=Southwick2000>{{Cite journal| surname= Southwick | name= Frederick S.| year= 2000| title = Gelsolin and ADF/cofilin enhance the actin dynamics of motile cells| pub-journal= Proceedings of the National Academy of Sciences of the United States of America| volume= 97| number= 13| page = 6936| doi = 10.1073/pnas.97.13.6936| url = http://www.pnas.org/cgi/content/full/pnas;97/13/6936| pmid = 10860951}}</ref>

Other proteins that bind with actin cover the ends of F-actin in order to stabilize them, but they are unable to break them. Examples of this type of protein are [[CapZ]] (that binds the (+) ends depending on a cell’s levels of [[calcium|Ca<sup>2+</sup>]]/[[calmodulin]]. These levels depend on the cell’s internal and external signals and are involved in the regulation of its biological functions).<ref name=Caldwell1989>{{Cite journal | surname1= Caldwell | name1= J.E. | surname2= Heiss | name2= S.G. | surname3= Mermall | name3= V. | surname4= Cooper | name4= J.A. | year= 1989 | title = Effects of CapZ, an actin-capping protein of muscle, on the polymerization of actin | pub-journal= Biochemistry | volume= 28 | number= 21 | page = 8506–8514 | doi = 10.1021/bi00447a036 | url = http://pubs.acs.org/doi/abs/10.1021/bi00447a036 | format= w}}</ref> Another example is [[tropomodulin]] (that binds to the (-) end). Tropomodulin basically acts to stabilize the F-actin present in the [[myofibril]]s present in [[muscle]] [sarcomere]]s, which are structures characterized by their great stability.<ref name=Weber1994>{{Cite journal | surname1= Weber | name1= A. | surname2= Pennise | name2= C.R. | surname3= Babcock | name3= G.G. | surname4= Fowler | name4= V.M. | year= 1994 | title = Tropomodulin caps the pointed ends of actin filaments | pub-journal= The Journal of Cell Biology | volume= 127 | number= 6 | page = 1627–1635 | doi = 10.1083/jcb.127.6.1627 | url = http://www.jcb.org/cgi/reprint/127/6/1627.pdf | pmid = 7798317}}</ref>

[[Image:Arp2 3 complex.png|thumb|Atomic structure of Arp2/3.<ref>Robinson RC, Turbedsky K, Kaiser DA, Marchand JB, Higgs HN, Choe S, Pollard TD. (2001) [http://www.ncbi.nlm.nih.gov/pubmed/11721045 Crystal structure of Arp2/3 complex]. Science 294(5547):1679-84.</ref> Each colour corresponds to a subunit: Arp3, orange; Arp2, sea blue (subunits 1 and 2 are not shown); p40, green; p34, light blue; p20, dark blue; p21, magenta; p16, yellow.]]

The [[Arp2/3]] complex is widely found in all [[Eukaryote|eukaryotic]] organisms.<ref>{{Cite journal|author=Mullins, R. D.|co-authors=Pollard, T.D.|title= Structure and function of the Arp2/3 complex|revista=Current Opinion in Structural Biology|volume=9|number=2|fecha=April 1999|pages=244–249|publisher=Elsevier|doi=10.1016/S0959-440X(99)80034-7|url=http://www.ingentaconnect.com/content/els/0959440x/1999/00000009/00000002/art80034?token=00431ce3643f038fde4775686f2357275c277b422c40465d483f2544446e7b6dea2|fechaaceso=2007-10-03}}</ref> It is comprised of seven subunits, some of which possess a [[topology]] that is clearly related to their biological function: two of the subunits, «ARP2» and «ARP3», have a structure similar to that of actin monomers. This homology allows both units to act as [[nucleation|nucleation agents]] in the polymerization of G-actin and F-actin. This complex is also required in more complicated processes such as in establishing [[dendrite|dendritic]] structures and also in [[anastomosis]] (the reconnection of two branching structures that had previously been joined, such as in blood vessels).<ref name="Machesky and Gould">{{Cite journal|author=Laura M Machesky|coauthors=Kathleen L Gould|title= The Arp2/3 complex: a multifunctional actin organizer|revista=Current Opinion in Cell Biology|volume=11|number=1|date=February 1999|pages=117-121|doi=doi:10.1016/S0955-0674(99)80014-3 |url=http://www.sciencedirect.com/science?_ob=ArticleURL&_udi=B6VRW-3W8SD3R-G&_user=10&_rdoc=1&_fmt=&_orig=search&_sort=d&view=c&_acct=C000050221&_version=1&_urlVersion=0&_userid=10&md5=c399e9d447df11d385cbefe9b44056d7}}</ref>

=== Chemical inhibitors ===

[[Image:Phalloidin.png|thumb|Chemical structure of [[phalloidin]].]]

There are a number of [[toxin]]s that interfere with actin’s dynamics, either by preventing it from polymerizing ([[latrunculin]] and [[cytochalasin D]]) or by stabilizing it ([[phalloidin]]):
* Latrunculin is a toxin produced by [[sponge]]s, it binds to G-actin preventing it from binding with microfilaments.<ref name=Morton2000>{{Cite journal | surname1= Morton | name1= W.M. | surname2= Ayscough | name2= K.R. | surname3= McLaughlin | name3= P.J. | year= 2000 | title = Latrunculin alters the actin-monomer subunit interface to prevent polymerization | pub-journal= Nature Cell Biology | volume= 2 | number= 6 | page = 376–378 | doi = 10.1038/35014075 | url = http://www.era.lib.ed.ac.uk/retrieve/1796/Mclaughlin.pdf}}</ref>
* Cytocalasin D, is an [[alkaloid]] produced by [[fungi]], that binds to the (+) end of F-actin preventing the addition of new monomers.<ref name=Cooper1987>{{Cite journal | surname= Cooper | name= J.A. | year= 1987 | title = Effects of cytochalasin and phalloidin on actin | pub-journal= The Journal of Cell Biology | volume= 105 | number= 4 | page = 1473–1478 | doi = 10.1083/jcb.105.4.1473 | url = http://www.jcb.org/cgi/reprint/105/4/1473.pdf}}</ref> Cytocalasin D has been found to disrupt actin’s dynamics in protein [[p53]], which has been found to affect the protein’s activity in [[animal]]s<ref name=Rubtsova1998>{{Cite journal | surname1= Rubtsova | name1= S.N. | surname2= Kondratov | name2= R.V. | surname3= Kopnin | name3= P.B. | surname4= Chumakov | name4= P.M. | surname5= Kopnin | name5= B.P. | surname6= Vasiliev | name6= J.M. | year= 1998 | title = Disruption of actin microfilaments by cytochalasin D leads to activation of p53 | pub-journal= FEBS Letters | volume= 430 | number= 3 | page = 353–357 | doi = 10.1016/S0014-5793(98)00692-9 | url = http://linkinghub.elsevier.com/retrieve/pii/S0014579398006929}}</ref> and [[gravitropism|gravitropic]] effects in [[plant]]s.<ref name=Staves1997>{{Cite journal | surname1= Staves | name1= M.P. | surname2= Wayne | name2= R. | surname3= Leopold | name3= A.C. | year= 1997 | title = Cytochalasin D does not inhibit gravitropism in roots | pub-journal= American Journal of Botany | volume= 84 | number= 11 | page = 1530–1530 | doi = 10.2307/2446614 | url = http://www.amjbot.org/cgi/reprint/84/11/1530.pdf}}</ref>
* Phalloidin, is a toxin that has been isolated from the death cap mushroom ''[[Amanita phalloides]]'', it binds to the interface between adjacent actin monomers in the F-actin polymer, preventing its depolymerization.<ref name=Cooper1987 />

== Functions and location ==

Actin forms [[microfilament]]s that are typically one of the most dynamic of the three subclasses of the eukaryotic [[cytoskeleton]]. This gives actin major functions in cells:
* Formation of [[microfilament]]s to give mechanical support to cells, and provide trafficking routes through the cytoplasm to aid signal transduction
* [[Cell motility]] in cells which undergo [[amoeboid]] motion using [[pseudopods]] (see [[Microfilament#Microfilament-based motility by actoclampin molecular motors|actoclampin molecular motors]]) and [[phagocytosis]], for example of bacteria by [[macrophage]]s
* In [[metazoa]]n [[muscle]] cells, to be the scaffold on which [[myosin]] proteins generate force to support muscle contraction
* In [[metazoa]]n [[muscle]] cells, to be the scaffold on which [[myosin]] proteins generate force to support muscle contraction
* In nonmuscle cells, to be a track for cargo transport myosins (nonconventional myosins) such as myosin V and VI. Nonconventional myosins use ATP hydrolysis to transport cargo, such as [[Vesicle (biology)|vesicles]] and organelles, in a directed fashion much faster than diffusion. Myosin V walks towards the barbed end of actin filaments, while myosin VI walks toward the pointed end. Most actin filaments are arranged with the barbed end toward the cellular membrane and the pointed end toward the cellular interior. This arrangement allows myosin V to be an effective motor for export of cargos, and myosin VI to be an effective motor for import.
* In non-muscle cells, to be a track for cargo transport myosins (nonconventional myosins) such as myosin V and VI. Nonconventional myosins use ATP hydrolysis to transport cargo, such as [[Vesicle (biology)|vesicles]] and organelles, in a directed fashion much faster than diffusion. Myosin V walks towards the barbed end of actin filaments, while myosin VI walks toward the pointed end. Most actin filaments are arranged with the barbed end toward the cellular membrane and the pointed end toward the cellular interior. This arrangement allows myosin V to be an effective motor for the export of cargos, and myosin VI to be an effective motor for import.
The actin protein is found in both the [[cytoplasm]] and the [[cell nucleus]].<ref name=Grummt2006 /> Its location is regulated by cell membrane [[signal transduction]] pathways that integrate the stimuli that a cell receives stimulating the restructuring of the actin networks in response. In ''[[Dictyostelium]]'', [[phospholipase D] has been found to intervene in [[inositol phosphate]] pathways.<ref>{{Cite journal | surname= M.V. Wakelam | name= T.R. Insall | year= 2005 | title = Phospholipase D activity is essential for actin localization and actin-based motility in Dictyostelium | pub-journal= Biochemical Journal | volume= 389 | number= Pt 1 | page = 207 | doi = 10.1042/BJ20050085 | url = http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=1184553 | format= w}}</ref> Actin filaments are particularly stable and abundant in [[myocyte|muscle fibres]]. Within the [[sarcomere]] (the basic morphological and physiological unit of muscle fibres) actin is present in both the I and A bands; myosin is also present in the latter.<ref name="eckert">{{Cite book | surnames = Randall | name = D. | co-authors = Burggren, W. et French, K. | title = Eckert Fisiología animal| edition = 4ª | id = ISBN 84-486-0200-5}}</ref>


===Directionality===
=== Cytoskeleton ===
[[Image:MEF microfilaments.jpg|thumb|[[Fluorescence microscope|Fluorescence]] micrograph showing F-actin (in green) in rat [[fibroblast]]s.]]
The polarity of an actin filament can be determined by decorating the microfilament with myosin "S1" fragments, creating barbed (+) and pointed (-) ends on the filament. An S1 fragment is composed of the head and neck domains of myosin II. Under physiologic conditions, G-actin (the [[monomer]] form) is transformed to F-actin (the [[polymer]] form) by ATP, where the role of ATP is essential<ref>HISTOLOGI by Finn Geneser p. 105</ref>


{{Main | Microfilament}}
===Nucleation and polymerization===


Microfilaments are involved in the movement of all mobile cells, including non-muscular types, and drugs that disrupt F-actin organization (such as the [[cytochalasin]]s) affect the activity of these cells. Actin comprises 2% of the total amount of proteins in [[hepatocyte]]s, 10% in [[fibroblast]]s, 15% in [[amoeba]]s and up to 50-80% in activated [[platelet]]s.<ref name="Pujol">{{cite book|surnames= Pujol-Moix|title= Trombocitopenias
Actin polymerization and depolymerization is necessary in [[chemotaxis]] and [[cytokinesis]]. Nucleating factors are necessary to stimulate actin polymerization. One such nucleating factor is the Arp2/3 complex, which mimics a G-actin dimer to stimulate the nucleation of G-actin (or monomeric actin). The [[Arp2/3 complex]] binds to actin filaments at 70 degrees to form new actin branches off of existing actin filaments. Also, actin filaments themselves bind ATP, and hydrolysis of this ATP stimulates destabilization of the polymer.
|edition=2nd|url=http://books.google.es/books?id=l_X1vOPyyl4C&source=gbs_navlinks_s|year= 2001|publisher= [[Elsevier]], España|language= Spanish|isbn= 8481745952|pages=25}}</ref> There are a number of different types of actin with slightly different structures and functions. This means that α-actin is found exclusively in [[muscle fibre]]s, while types β and γ are found in other cells. In addition, as the latter types have a high turnover rate the majority of them are found outside permanent structures. This means that the microfilaments found in cells other than muscle cells are present in two forms:<ref name="paniagua" />
* Microfilament networks. [[Animal cell]]s commonly have a cell cortex under the [[cell membrane]] that contains a large number of microfilaments, which precludes the presence of [[organelle]]s. This network is connected with numerous [[Receptor (biochemistry)|receptor cell]]s that [[signal transduction|relay signals]] to the outside of a cell.
* Microfilament bundles. These extremely long microfilaments are located in networks and, in association with contractile proteins such as non-muscular [[myosin]], they are involved in the movement of substances at an intracellular level.


==== Yeasts ====
The growth of actin filaments can be regulated by [[thymosin]] and [[profilin]]. Thymosin binds to G-actin to buffer the polymerizing process, while profilin binds to G-actin to exchange [[Adenosine diphosphate|ADP]] for [[Adenosine triphosphate|ATP]], promoting the monomeric addition to the barbed, plus end.
Actin’s cytoskeleton is key to the processes of [[endocytosis]], [[cytokinesis]], determination of [[cell polarity]] and [[morphogenesis]] in [[yeast]]s. In addition to relying on actin these processes involve 20 or 30 associated proteins, which all have a high degree of evolutionary conservation, along with many signalling molecules. Together these elements allow a spatially and temporally modulated assembly that defines a cell’s response to both internal and external stimuli.<ref name=Moseley2006>{{Cite journal | surname1= Moseley | name1= James B. | surname2= Goode | name2= Bruce L. | year= 2006 | title = The Yeast Actin Cytoskeleton: from Cellular Function to Biochemical Mechanism | pub-journal= Microbiology and Molecular Biology Reviews | volume= 70 | number= 3 | page = 605–645 | doi = 10.1128/MMBR.00013-06 | url = http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=1594590 | pmid = 16959963}}</ref>

Yeasts contain three main elements that are associated with actin: patches, cables and rings that, despite being present for long periods of time, are subject to a dynamic equilibrium due to continual polymerization and depolymerization. They possess a number of accessory proteins including ADF/cofilin, which has a molecular weight of 16kDa and is coded for by a single gene, called ''COF1''; Aip1, a cofilin cofactor that promotes the disassembly of microfilaments; Srv2/CAP, a process regulator related to [[adenylate cyclase]] proteins; a profilin with a molecular weight of approximately 14 kDa that is associated with actin monomers; and twinfilin, a 40 kDa protein involved in the organization of patches.<ref name=Moseley2006 />

==== Plants ====
Plant [[genome]] studies have revealed the existence of protein isovariants within the actin family of genes. Within ''[[Arabidopsis thaliana]]'', a [[dicotyledon]] used as a [[model organism]], there are ten types of actin, nine types of α-tubulins, six β-tubulins, six profilins and dozens of myosins. This diversity is explained by the evolutionary necessity of possessing variants that slightly differ in their temporal and spatial expression. The majority of these proteins were jointly expressed in the [[Tissue (biology)|tissue]] analysed. Actin networks are distributed throughout the cytoplasm of cells that have been cultivated ''[[in vitro]]''. There is a concentration of the network around the nucleus that is connected via spokes to the cellular cortex, this network is highly dynamic, with a continuous polymerization and depolymerization.<ref name=Meagher1999>{{Cite journal | surname1= Meagher | name1= R.B. | surname2= McKinney | name2= E.C. | surname3= Kandasamy | name3= M.K. | year= 1999 | title = Isovariant Dynamics Expand and Buffer the Responses of Complex Systems: The Diverse Plant Actin Gene Family | pub-journal= The Plant Cell Online | volume= 11 | number= 6 | page = 995–1006 | url = http://www.plantcell.org/cgi/content/full/11/6/995}}</ref>

[[Image:PDB 1unc EBI.jpg|thumb|left|[[Protein structure|Structure]] of the C-terminal subdomain of [[villin]], a protein capable of splitting microfilaments.<ref>[http://www.ebi.ac.uk/pdbe-srv/view/entry/1unc/summary Entrada PDBe para 1unc]. [[European Bioinformatics Institute|EBI]].</ref>]]

Even though the majority of plant cells have a [[Plant cell wall|cell wall]] that defines their morphology and impedes their movement, their microfilaments can generate sufficient force to achieve a number of cellular activities, such as, the cytoplasmic currents generated by the microfilaments and myosin. Actin is also involved in the movement of organelles and in cellular morphogenesis, which involve [[cell division]] as well as the elongation and differentiation of the cell.<ref name=Higaki2007>{{Cite journal | surname1= Higaki | name1= T. | surname2= Sano | name2= T. | surname3= Hasezawa | name3= S. | year= 2007 | title = Actin microfilament dynamics and actin side-binding proteins in plants | pub-journal= Current Opinion in Plant Biology | volume= 10 | number= 6 | page = 549–556 | doi = 10.1016/j.pbi.2007.08.012 | url = http://lmbiologia.campusnet.unito.it/didattica/att/a29a.1275.file.pdf | format= PDF}}</ref>

The most notable proteins associated with the actin cytoskeleton in plants include:<ref name=Higaki2007 /> [[villin]], which belongs to the same family as [[gelsolin]]/severin and is able to cut microfilaments and bind actin monomers in the presence of calcium cations; [[fimbrin]], which is able to recognize and unite actin monomers and which is involved in the formation of networks (by a different regulation process from that of animals and yeasts);<ref name=Kovar2000>{{Cite journal | surname1= Kovar | name1= D.R. | surname2= Staiger | name2= C.J. | surname3= Weaver | name3= E.A. | surname4= McCurdy | name4= D.W. | year= 2000 | title = AtFim1 is an actin filament crosslinking protein from Arabidopsis thaliana | pub-journal= The Plant Journal | volume= 24 | number= 5 | page = 625–636 | doi = 10.1046/j.1365-313x.2000.00907.x}}</ref> [[formin]]s, which are able to act as an F-actin polymerization nucleating agent; [[myosin]], a typical molecular motor that is specific to eukaryotes and which in ''Arabidopsis thaliana'' is coded for by 17 genes in two distinct classes; CHUP1, which can bind actin and is implicated in the spatial distribution of [[chloroplast]]s in the cell; KAM1/MUR3 that define the morphology of the [[Golgi apparatus]] as well as the composition of [[xyloglucan]]s in the cell wall; NtWLIM1, which facilitates the emergence of actin cell structures; and ERD10, which is involved in the association of organelles within [[cell membrane|membranes]] and microfilaments and which seems to play a role that is involved in an organism’s reaction to [[Stress (biology)|stress]].

===Nuclear actin===
Actin is essential for [[transcription (genetics)|transcription]] from RNA polymerases [[RNA polymerase I|Pol I]], [[RNA polymerase II|Pol II]] and [[RNA polymerase III|Pol III]]. In Pol I transcription, actin and myosin ([[MYO1C]], which binds DNA) act as a [[molecular motor]]. For Pol II transcription, β-actin is needed for the formation of the preinitiation complex. Pol III contains β-actin as a subunit. Actin can also be a component of chromatin remodelling complexes as well as pre-mRNP particles (that is, precursor [[messenger RNA]] bundled in proteins), and is involved in [[nuclear pore|nuclear export]] of RNAs and proteins.<ref>{{cite journal |author=Zheng B, Han M, Bernier M, Wen JK |title=Nuclear actin and actin-binding proteins in the regulation of transcription and gene expression |journal=FEBS J. |volume=276 |issue=10 |pages=2669–85 |year=2009 |month=May |pmid=19459931 |pmc=2978034 |doi=10.1111/j.1742-4658.2009.06986.x}}</ref>


===Microfilaments===
Individual [[Protein subunit|subunit]]s of [[microfilaments]] are known as [[Globular protein|globular]] actin (G-actin). G-actin subunits assemble into long filamentous [[Biopolymer|polymer]]s called F-actin. Two parallel F-actin strands must rotate 166 degrees to layer correctly on top of each other. This creates the double helix structure of the microfilaments of the cytoskeleton. Microfilaments measure approximately 7 [[nanometer|nm]] in diameter with a loop of the helix repeating every 37&nbsp;nm.


=== Muscular contraction ===
===Actomyosin filaments===
{{Main |Muscular contraction}}
[[Image:Sarcomere es.svg|thumb|The structure of a [[sarcomere]], the basic morphological and functional unit of the skeletal muscles that contains actin.]]
====Outline of a muscle contraction====
In [[muscle]], actin is the major component of ''thin filaments'', which, together with the [[motor protein]] [[myosin]] (which forms ''thick filaments''), are arranged into actomyosin [[myofibril]]s. These fibrils comprise the mechanism of [[muscle contraction]]. Using the hydrolysis of [[Adenosine triphosphate|ATP]] for energy, myosin heads undergo a cycle during which they attach to thin filaments, exert a tension, and then, depending on the load, perform a power stroke that causes the thin filaments to slide past, shortening the muscle.
In [[muscle]], actin is the major component of ''thin filaments'', which, together with the [[motor protein]] [[myosin]] (which forms ''thick filaments''), are arranged into actomyosin [[myofibril]]s. These fibrils comprise the mechanism of [[muscle contraction]]. Using the hydrolysis of [[Adenosine triphosphate|ATP]] for energy, myosin heads undergo a cycle during which they attach to thin filaments, exert a tension, and then, depending on the load, perform a power stroke that causes the thin filaments to slide past, shortening the muscle.


In contractile bundles, the actin-bundling protein alpha-[[actinin]] separates each thin filament by ~35&nbsp;nm. This increase in distance allows thick filaments to fit in between and interact, enabling deformation or contraction. In deformation, one end of myosin is bound to the [[plasma membrane]], while the other end "walks" toward the plus end of the actin filament. This pulls the membrane into a different shape relative to the [[cell cortex]]. For contraction, the myosin molecule is usually bound to two separate filaments and both ends simultaneously "walk" toward their filament's plus end, sliding the actin filaments closer to each other. This results in the shortening, or contraction, of the actin bundle (but not the filament). This mechanism is responsible for muscle contraction and [[cytokinesis]], the division of one cell into two.
In contractile bundles, the actin-bundling protein alpha-[[actinin]] separates each thin filament by ~35&nbsp;nm. This increase in distance allows thick filaments to fit in between and interact, enabling deformation or contraction. In deformation, one end of myosin is bound to the [[plasma membrane]], while the other end "walks" toward the plus end of the actin filament. This pulls the membrane into a different shape relative to the [[cell cortex]]. For contraction, the myosin molecule is usually bound to two separate filaments and both ends simultaneously "walk" toward their filament's plus end, sliding the actin filaments closer to each other. This results in the shortening, or contraction, of the actin bundle (but not the filament). This mechanism is responsible for muscle contraction and [[cytokinesis]], the division of one cell into two.
====Actin’s role in muscle contraction====
The helical F-actin filament found in muscles also contains a [[tropomyosin]] molecule, which is a 40 [[nanometre]]s long protein that is wrapped around the F-actin helix. During the resting phase the tropomyosin covers the actin’s active sites so that the actin-myosin interaction cannot take place and produce muscular contraction (the interaction gives rise to a movement between the two proteins that, because it is repeated many times, produces a contraction). There are other protein molecules bound to the tropomyosin thread, these include the [[troponin]]s that have three polymers: [[troponin I]], [[troponin T]] and [[troponin C]].<ref name=fisiologia /> Tropomyosin’s regulatory function depends on its interaction with troponin in the presence of Ca<sup>2+</sup> ions.<ref>Antoni Bayés de Luna, VV Staff, José López-Sendón, Fause Attie, Eduardo Alegría Ezquerra [http://books.google.es/books?id=OEFvw6RRgBoC Cardiología Clínica] (in Spanish). Published by Elsevier España, 2002; page 19. ISBN 84-458-1179-7</ref>


Both actin and [[myosin]] are involved in [[muscle]] contraction and relaxation and they make up 90% of muscle protein.<ref name=bioq>John W. Baynes, Marek H. Dominiczak [http://books.google.es/books?id=OCWP08sZok4C Bioquímica médica] Published by [[Elsevier]] Spain, 2007; page 268. ISBN 84-8174-866-8</ref> The overall process is initiated by an external signal, typically through an [[action potential]] stimulating the muscle, which contains specialized cells whose interiors are rich in actin and myosin filaments. The contraction-relaxation cycle comprises the following steps:<ref name=autogenerated2>{{cite book | surnames = Randall | name = D. | co-authors = Burggren, W. et French, K. | title = Eckert Fisiología animal| edition = 4ª | id = ISBN 84-486-0200-5}}</ref>
===Nuclear actin===
# Depolarization of the [[sarcolemma]] and transmission of an action potential through the [[T-tubule]]s.
Actin is essential for [[transcription (genetics)|transcription]] from RNA polymerases [[RNA polymerase I|I]], [[RNA polymerase II|II]] and [[RNA polymerase III|III]]. In Pol I transcription, actin and myosin ([[MYO1C]], which binds DNA) act as a [[molecular motor]]. For Pol II transcription, β-actin is needed for the formation of the preinitiation complex. Pol III contains β-actin as a subunit. Actin can also be a component of chromatin remodeling complexes as well as pre-mRNP particles (that is, precursor [[messenger RNA]] bundled in proteins), and is involved in [[nuclear pore|nuclear export]] of RNAs and proteins.<ref>{{cite journal |author=Zheng B, Han M, Bernier M, Wen JK |title=Nuclear actin and actin-binding proteins in the regulation of transcription and gene expression |journal=FEBS J. |volume=276 |issue=10 |pages=2669–85 |year=2009 |month=May |pmid=19459931 |pmc=2978034 |doi=10.1111/j.1742-4658.2009.06986.x}}</ref>
# Opening of the [[sarcoplasmic reticulum]]’s [[calcium|Ca<sup>2+</sup>]] channels.
# Increase in [[cytosol|cytosolic]] Ca<sup>2+</sup> concentrations and the interaction of these cations with troponin causing a conformational change in its [[protein structure|structure]]. This in turn alters the structure of tropomyosin, which covers actin’s active site, allowing the formation of myosin-actin cross-links (the latter being present as thin filaments).<ref name=fisiologia />
# Movement of myosin heads over the thin filaments, this can either involve ATP or be independent of ATP. The former mechanism, mediated by [[ATPase]] activity in the myosin heads, causes the movement of the actin filaments towards the [[Sarcomere|Z-disc]].
# Ca<sup>2+</sup> capture by the sarcoplasmic reticulum, causing a new conformational change in tropomyosin that inhibits the actin-myosin interaction.<ref name=bioq />

=== Other biological processes ===
The traditional image of actin’s function relates it to the maintenance of the cytoskeleton and, therefore, the organization and movement of organelles, as well as the determination of a cell’s shape.<ref name="paniagua">{{cite book| author = Paniagua, R.; Nistal, M.; Sesma, P.; Álvarez-Uría, M.; Fraile, B.; Anadón, R. and José Sáez, F. | title = Citología e histología vegetal y animal | year = 2002 | publisher = McGraw-Hill Interamericana de España, S.A.U. | Language = Spanish | id = ISBN 84-486-0436-9}}</ref> However, actin has a wider role in eukaryotic cell physiology, in addition to similar functions in [[prokaryote]]s.

* [[Cytokinesis]]. [[Cell division]] in animal cells and yeasts normally involves the separation of the parent cell into two daughter cells through the constriction of the central circumference. This process involves a constricting ring composed of actin, myosin and [[actinin|α-actinin]].<ref name=Fujiwara1978>{{Cite journal | surname1= Fujiwara | name1= K. | surname2= Porter | name2= M.E. | surname3= Pollard | name3= T.D. | year= 1978 | title = Alpha-actinin localization in the cleavage furrow during cytokinesis | pub-journal= The Journal of Cell Biology | volume= 79 | number= 1 | page = 268–275 | doi = 10.1083/jcb.79.1.268 | url = http://www.jcb.org/cgi/reprint/79/1/268.pdf | pmid = 359574}}</ref> In the "fission yeast” ''[[Schizosaccharomyces pombe]]'', actin is actively formed in the constricting ring with the participation of [[Arp2/3|Arp3]], the [[formin]] Cdc12, [[profilin]] and [[WASp]], along with preformed microfilaments. Once the ring has been constructed the structure is maintained by a continual assembly and disassembly that, aided by the [[Arp2/3]] complex and formins, is key to one of the central processes of cytokinesis.<ref name=Pelham2002>{{Cite journal | surname1= Pelham | name1= R.J. | surname2= Chang | name2= F. | year= 2002 | title = Actin dynamics in the contractile ring during cytokinesis in fission yeast | pub-journal= Nature | volume= 419 | number= 6902 | page = 82–86 | url = http://adsabs.harvard.edu/abs/2002Natur.419...82P}}</ref> The totality of the contractile ring, the [[spindle apparatus]], [[microtubule]]s and the dense peripheral material is called the «Fleming body» or «intermediate body».<ref name="paniagua" />

* [[Apoptosis]]. During [[programmed cell death]] the ICE/ced-3 family of proteases (one of the interleukin-1β-converter proteases) degrade actin into two fragments “in vivo”, one of the fragments is 15 kDa and the other 31 kDa. This represents one of the mechanisms involved in destroying cell viability that form the basis of apoptosis.<ref name=Mashima1997>{{Cite journal | surname1= Mashima | name1= T. | surname2= Naito | name2= M. | surname3= Noguchi | name3= K. | surname4= Miller | name4= D.K. | surname5= Nicholson | name5= D.W. | surname6= Tsuruo | name6= T. | year= 1997 | title = Actin cleavage by CPP-32/apopain during the development of apoptosis | pub-journal= Oncogene | volume= 14 | number= 9 | page = 1007–1012 | doi = 10.1038/sj.onc.1200919 | url = http://www.nature.com/onc/journal/v14/n9/abs/1200919a.html | format= w}}</ref> The protease [[calpain]] has also been shown to be involved in this type of cell destruction;<ref name=Wang2000>{{Cite journal | surname= Wang | name= K.K.W. | year= 2000 | title = Calpain and caspase: can you tell the difference? | pub-journal= Trends in Neurosciences | volume= 23 | number= 1 | page = 20–26 | doi = 10.1016/S0166-2236(99)01479-4 | url = http://www.mbi.ufl.edu/ctbis/wang/Wang71TINS2000.pdf | format= w}}</ref> just as the use of calpain inhibitors has been shown to decrease actin proteolysis and the degradation of [[DNA]] (another of the characteristic elements of apoptosis).<ref name=Villa1998>{{Cite journal | surname= Villa | name= P.G. | year= 1998 |title = Calpain inhibitors, but not caspase inhibitors, prevent actin proteolysis and DNA fragmentation during apoptosis| pub-journal= Journal of Cell Science | volume= 111 | number= 6 | page = 713–722 | url = http://jcs.biologists.org/cgi/reprint/111/6/713.pdf}}</ref> On the other hand, the [[stress]]-induced triggering of apoptosis causes the reorganization of the actin cytoskeleton (which also involves its polymerization), giving rise to structures called [[stress fibres]]; this is activated by the [[MAPK/ERK pathway|MAP kinase]] pathway.<ref name=Huot1998>{{Cite journal | surname1= Huot | name1= J. | surname2= Houle | name2= F. | surname3= Rousseau | name3= S. | surname4= Deschesnes | name4= R.G. | surname5= Shah | name5= G.M. | surname6= Landry | name6= J. | year= 1998 | title = SAPK2/p38-dependent F-Actin Reorganization Regulates Early Membrane Blebbing during Stress-induced Apoptosis | pub-journal= The Journal of Cell Biology | volume= 143 | number= 5 | page = 1361–1373 | url = http://www.jcb.org/cgi/content/full/143/5/1361}}</ref>

[[Image:Cellular tight junction keys.svg|thumb|Diagram of a ''[[zonula occludens]]'' or tight junction, a structure that joins the [[epithelium]] of two cells. Actin is one of the anchoring elements shown in green.]]

* [[Cellular adhesion]] and [[Developmental biology|development]]. The adhesion between cells is a characteristic of [[multicellular organisms]] that enables [[Tissue (biology)|tissue]] specialization and therefore increases cell complexity. Adhesion of cell [[epithelium|epithelia]] involves the actin cytoskeleton in each of the joined cells as well as [[cadherin]]s acting as extracellular elements with the connection between the two mediated by [[catenin]]s.<ref name=Adams1996>{{Cite journal | surname1= Adams | name1= C.L. | surname2= Nelson | name2= W.J. | surname3= Smith | name3= S.J. | year= 1996 | title = Quantitative analysis of cadherin-catenin-actin reorganization during development of cell-cell adhesion | pub-journal= The Journal of Cell Biology | volume= 135 | number= 6 | page = 1899–1911 | doi = 10.1083/jcb.135.6.1899 | url = http://www.jcb.org/cgi/reprint/135/6/1899.pdf | pmid = 8991100 | format= w}}</ref> Interfering in actin dynamics has repercussions for an organism’s development, in fact actin is such a crucial element that systems of redundant [[gene]]s are available. For example, if the [[α-actinin]] or [[gelation factor]] gene has been removed in ''[[Dictyostelium]]'' individuals do not show an anomalous [[phenotype]] possibly due to the fact that each of the proteins can perform the function of the other. However, the development of [[Mutation|double mutations]] that lack both gene types is affected.<ref name=Witke1992>{{Cite journal | surname1= Witke | name1= W. | surname2= Schleicher | name2= M. | surname3= Noegel | name3= A.A. | year= 1992 | title = Redundancy in the microfilament system: abnormal development of Dictyostelium cells lacking two F-actin cross-linking proteins | pub-journal= Cell | volume= 68 | number= 1 | page = 53–62 | url = http://www.ncbi.nlm.nih.gov/pubmed/1732064}}</ref>

* [[Gene expression]] modulation. Actin’s state of polymerization affects the pattern of [[gene expression]]. In [[1997]], it was discovered that cytocalasin D-mediated depolymerization in [[Schwann cell]]s causes a specific pattern of expression for the genes involved in the [[myelinization]] of this type of [[Neuron|nerve cell]].<ref name=Fernandez-valle1997>{{Cite journal | surname1= Fernandez-valle | name1= C. | surname2= Gorman | name2= D. | surname3= Gomez | name3= A.M. | surname4= Bunge | name4= M.B. | year= 1997 | title = Actin Plays a Role in Both Changes in Cell Shape and Gene- Expression Associated with Schwann Cell Myelination | pub-journal= Journal of Neuroscience | volume= 17 | number= 1 | page = 241–250 | url = http://www.jneurosci.org/cgi/content/full/17/1/241}}</ref> F-actin has been shown to modify the [[transcriptome]] in some of the life stages of unicellular organisms, such as the fungus ''[[Candida albicans]]''.<ref name=Wolyniak2007>{{Cite journal | surname1= Wolyniak | name1= Michael J. | surname2= Sundstrom | name2= Paula | year= 2007 | title = Role of Actin Cytoskeletal Dynamics in Activation of the Cyclic AMP Pathway and HWP1 Gene Expression in Candida albicans | pub-journal= Eukaryotic Cell | volume= 6 | number= 10 | page = 1824–1840 | doi = 10.1128/EC.00188-07 | url = http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2043390 | pmid = 17715368}}</ref> In addition, proteins that are similar to actin play a regulatory role during [[spermatogenesis]] in [[Muridae|mice]]<ref name=Tanaka2003>{{Cite journal | surname1= Tanaka | name1= Hiromitsu | surname2= Iguchi | name2= Naoko | surname3= Egydio De Carvalho | name3= Carlos | surname4= Tadokoro | name4= Yuko | surname5= Yomogida | name5= Kentaro | surname6= Nishimune | name6= Yoshitake | year= 2003 | title = Novel Actin-Like Proteins T-ACTIN 1 and T-ACTIN 2 Are Differentially Expressed in the Cytoplasm and Nucleus of Mouse Haploid Germ Cells | pub-journal= Biology of Reproduction | volume= 69 | number= 2 | page = 475–482 | doi = 10.1095/biolreprod.103.015867 | url = http://www.biolreprod.org/cgi/reprint/69/2/475.pdf | pmid = 12672658}}</ref> and, in yeasts, actin-like proteins are thought to play a role in the regulation of [[Epigenetics|gene expression]].<ref name=Jiang1996>{{Cite journal | surname1= Jiang | name1= Y.W. | surname2= Stillman | name2= D.J. | year= 1996 | title = Epigenetic effects on yeast transcription caused by mutations in an actin-related protein present in the nucleus | pub-journal= Genes & Development | volume= 10 | number= 5 | page = 604–619 | doi = 10.1101/gad.10.5.604 | url = http://www.genesdev.org/cgi/reprint/10/5/604.pdf | format= w}}</ref> In fact, actin is capable of acting as a transcription initiator when it reacts with a type of nuclear myosin that interacts with [[RNA polymerase]]s and other enzymes involved in the transcription process.<ref name=Grummt2006>{{Cite journal | surname= Grummt | name= I | year= 2006 | title = Actin and myosin as transcription factors | pub-journal= Current opinion in genetics & development | volume= 16 | number= 2 | page = 191–196 | doi = 10.1016/j.gde.2006.02.001 | url = http://linkinghub.elsevier.com/retrieve/pii/S0959437X06000232 | format= w}}</ref>

* [[Stereocilia]] dynamics. Some cells develop fine filliform outgrowths on their surface that have a [[Somatosensory system|mechanosensory]] function. For example, this type of organelle is present in the [[Organ of Corti]], which is located in the [[ear]] . The main characteristic of these structures is that their length can be modified.<ref>{{Cite journal | title = Dynamic length regulation of sensory stereocilia | url = http://linkinghub.elsevier.com/retrieve/pii/S1084952108000451 | year= 2008 | pub-journal= Seminars in Cell and Developmental Biology | page = 502–510 | volume= 19 | number= 6 | surname1= Manor | name1= U. | surname2= Kachar| name2= B. | access date= 20 June 2009}}</ref> The molecular architecture of the stereocilia includes a [[paracrystalline]] actin core in dynamic equilibrium with the monomers present in the adjacent cytosol. Type VI and VIIa myosins are present throughout this core, while myosin XVa is present in its extremities in quantities that are proportional to the length of the stereocilia.<ref>{{Cite journal | title = An actin molecular treadmill and myosins maintain stereocilia functional architecture and self- | url = http://jcb.rupress.org/cgi/content/full/164/6/887 | year= 2004 | pub-journal= Journal of Cell Biology | page = 887–897 | volume= 164 | number= 6 | surname1= Rzadzinska| name1= A.K. | surname2= Schneider | name2= M.E. | surname3= Davies | name3= C. | surname4= Riordan| name4= G.P. | surname5= Kachar | name5= B. | access date= 20 June 2009}}</ref>

== Molecular pathology ==
The majority of [[mammal]]s posses six different actin [[gene]]s. Of these, two code for the [[cytoskeleton]] (''[[ACTB]]'' and ''[[ACTG1]]'') while the other four are involved in [[skeletal striated muscle]] (''[[ACTA1]]''), [[smooth muscle tissue]] (''[[ACTA2]]''), [[Intestine|intestinal]] muscles (''[[ACTG2]]'') and [[cardiac muscle]] (''[[ACTC1]]''). In this way the actin in the cytoskeleton is involved in the [[Pathogenesis|pathogenic]] mechanisms of many [[Pathogen|infectious agents]], including [[HIV]]. The vast majority of the [[mutation]]s that affect actin are point mutations that have a [[Dominance (genetics)|dominant effect]], with the exception of six mutations involved in [[nemaline myopathy]]. This is because in many cases the mutant of the actin monomer acts as a “cap” by preventing the elongation of F-actin.<ref name="dosremedios" />

=== Pathology associated with ''ACTA1'' ===
''[[ACTA1]]'' is the gene that codes for the α-[[isoform]] of actin that is predominant in human [[skeletal striated muscle]]s, although it is also expressed in heart muscle and in the [[Thyroid|thyroid gland]].<ref>{{cite journal| author =Su AI, Wiltshire T, Batalov S, Lapp H, Ching KA, Block D, Zhang J, Soden R, Hayakawa M, Kreiman G, Cooke MP, Walker JR, Hogenesch JB| title =A gene atlas of the mouse and human protein-encoding transcriptomes | year =2004 | journal = Proc Natl Acad Sci U S A. | volume =101 | number =16 | id = PMID 15075390| url =http://www.pnas.org/content/101/16/6062.full}}</ref> Its [[DNA sequence]] consists of seven [[exon]]s that produce five known [[Transcription (genetics)|transcripts]].<ref name="uniprot">{{cite web|url =http://www.uniprot.org/uniprot/P68133 |title =Uniprot ACTA1 |access date =[[26 December]]|access year =2008 }}</ref> The majority of these consist of point mutations causing substitution of [[amino acid]]s. The mutations are in many cases associated with a [[phenotype]] that determines the severity and the course of the affliction.<ref name="dosremedios">{{cite book| author =Cristóbal G. Dos Remedios, Deepak Chhabra | title =Actin-binding Proteins and Disease | year =2008 | publisher = Springer | id =ISBN 0-387-71747-1}} See in [http://books.google.es/books?id=Y6xHVnH8iOIC&pg=PA23&dq=actinopathy&client=firefox-a#PPA18,M1 Google books]</ref><ref name="uniprot" />

[[Image:Nemaline rods.jpg|thumb|Giant [[Nemaline myopathy|nemaline rods]] produced by the [[transfection]] of a [[DNA sequence]] of ''[[ACTA1]]'', which is the carrier of a [[mutation]] responsible for nemaline myopathy.<ref name="Bathe" />]]

The mutation alters the structure and function of skeletal muscles producing one of three forms of [[myopathy]]: type 3 [[nemaline myopathy]], [[Congenital myopathy|congenital myopathy with an excess of thin myofilaments]] (CM) and [[Congenital myopathy#Congential fibre type disproportion|Congenital myopathy with fibre type disproportion]] (CMFTD). Mutations have also been found that produce [[Central core disease|“core” myopathies]]).<ref>{{cite journal| author =Kaindl AM, Rüschendorf F, Krause S, Goebel HH, Koehler K, Becker C, Pongratz D, Müller-Höcker J, Nürnberg P, Stoltenburg-Didinger G, Lochmüller H, Huebner A.| title =Missense mutations of ACTA1 cause dominant congenital myopathy with cores | year =2004 | journal = J Med Genet. | volume =41 | number =11 | id = PMID 15520409| url =}}</ref> Although their phenotypes are similar, in addition to typical nemaline myopathy some specialists distinguish another type of myopathy called actinic nemaline myopathy. In the former, clumps of actin form instead of the typical rods. It is important to state that a patient can show more than one of these [[phenotype]]s in a [[biopsy]].<ref name="sparrow">{{cite journal| surname =Sparrow | name =JC| co-authors= Nowak KJ, Durling HJ, Beggs AH, Wallgren-Pettersson C, Romero N, Nonaka I, Laing NG. | year=2003 |month=September | title =Muscle disease caused by mutations in the skeletal muscle alpha-actin gene (ACTA1) | journal =Neuromuscul Disord. | volume=13 | number=8 | pages=519-31| pmid=12921789}}</ref> The most common [[symptom]]s consist of a typical [[Face|facial]] morphology (myopathic faces), muscular weakness, a delay in motor development and respiratory difficulties. The course of the illness, its gravity and the age at which it appears are all variable and overlapping forms of myopathy are also found. A symptom of nemalinic myopathy is that “Nemaline rods” appear in differing places in Type 1 muscle fibres. These rods are non-[[pathognomonic]] structures that have a similar composition to the Z disks found in the [[sarcomere]].<ref name="revision">{{cite journal| author =North, Kathryn| title =Nemaline Myopathy | year =2002 | journal = Gene Reviews | id = PMID| url =http://www.ncbi.nlm.nih.gov/bookshelf/br.fcgi?book=gene&part=nem}}</ref>

The [[pathogenesis]] of this myopathy is very varied. Many mutations occur in the region of actin’s indentation near to its [[nucleotide]] binding sites, while others occur in Domain 2, or in the areas where interaction occurs with associated proteins. This goes some way to explain the great variety of clumps that form in these cases, such as Nemaline or Intranuclear Bodies or Zebra Bodies. <ref name="dosremedios" /> Changes in actin’s [[Protein folding|folding]] occur in nemaline myopathy as well as changes in its aggregation and there are also changes in the [[Gene expression|expression]] of other associated proteins. In some variants where intranuclear bodies are found the changes in the folding masks the [[Nuclear pore #Export of proteins|nucleus’s protein exportation signal]] so that the accumulation of actin's mutated form occurs in the [[cell nucleus]].<ref name="Ilkovski">{{cite journal| author =Ilkovski B, Nowak KJ, Domazetovska A, Maxwell AL, Clement S, Davies KE, Laing NG, North KN, Cooper ST.| title =Evidence for a dominant-negative effect in ACTA1 nemaline myopathy caused by abnormal folding, aggregation and altered polymerization of mutant actin isoforms | year =2004 | journal = Hum. Mol. Genet. | volume =13 | number =16 | id = PMID 15198992| url =http://hmg.oxfordjournals.org/cgi/content/full/13/16/1727}}</ref> On the other hand it appears that mutations to ''ACTA1'' that give rise to a CFTDM have a greater effect on sarcomeric function than on its structure.<ref name="Clarke">{{cite journal| author =Clarke NF, Ilkovski B, Cooper S, Valova VA, Robinson PJ, Nonaka I, Feng JJ, Marston S, North K| title =The pathogenesis of ACTA1-related congenital fibre type disproportion | year =2007 | journal = Ann Neurol | volume =61 | number =6 | id = PMID 17387733| url =}}</ref> Recent investigations have tried to understand this apparent paradox, which suggests there is no clear correlation between the number of rods and muscular weakness. It appears that some mutations are able to induce a greater [[apoptosis]] rate in type II muscular fibres.<ref name="Vandamme">{{cite journal| surname =Vandamme | name =D| co-authors=Lambert E, Waterschoot D, Cognard C, Vandekerckhove J, Ampe C, Constantin B, Rommelaere H.| year=2009 |month=July | title =alpha-Skeletal muscle actin nemaline myopathy mutants cause cell death in cultured muscle cells | journal =Biochim Biophys Acta. | volume=1793 | number=7 | pages=1259-71| pmid=19393268 }}</ref>

[[Image:Mutations in alpha actin.jpg|thumb|left|200px|Position of seven [[genetic mutation |mutations]] relevant to the various actinopathies related to ''[[ACTA1]]''.<ref name="Bathe">Friederike S Bathe, Heidi Rommelaere y Laura M Machesky (2007): [http://www.biomedcentral.com/1471-2121/8/2 Phenotypes of Myopathy-related Actin Mutants in differentiated C2C12 Myotubes]. BMC Cell Biology, 8:2. doi:10.1186/1471-2121-8-2</ref>]]

=== In smooth muscle ===
There are two isoforms that code for actins in the [[smooth muscle tissue]]:

''[[ACTG2]]'' codes for the largest actin isoform, which has nine [[exon]]s, one of which, the one located at the 5' end, is not [[Translation (biology)|translated]].<ref>{{cite journal| title =Structure, chromosome location, and expression of the human smooth muscle (enteric type) γ-actin gene: evolution of six human actin genes | year =1991 | journal = Mol Cell Biol. | volume =11 | number =6 | id = PMID 1710027| url =http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pubmed&pubmedid=1710027}}</ref> It is an γ-actin that is expressed in the enteric smooth muscle. No mutations to this gene have been found that correspond to pathologies, although [[DNA microarray|microarrays]] have shown that this protein is more often expressed in cases that are resistant to [[chemotherapy]] using [[cisplatin]].<ref>{{cite journal| author =Watson MB, Lind MJ, Smith L, Drew PJ, Cawkwell L.| title =Expression microarray analysis reveals genes associated with in vitro resistance to cisplatin in a cell line model | year =2007 | journal = Acta Oncol. | volume =46 | number =5 | id = PMID 17562441| url =}}</ref>

''[[ACTA2]]'' codes for an α-actin located in the smooth muscle, and also in vascular smooth muscle. It has been noted that the MYH11 mutation could be responsible for at least 14% of hereditary [[Aortic aneurism|thoracic aortic aneurisms]] particularly Type 6. This is because the mutated variant produces an incorrect filamentary assembly and a reduced capacity for vascular smooth muscle contraction. Degradation of the [[Aorta|aortic media]] has been recorded in these individuals, with areas of disorganization and [[hyperplasia]] as well as [[stenosis]] of the aorta’s [[vasa vasorum]].<ref>{{cite journal| author =Guo DC, Pannu H, Tran-Fadulu V, Papke CL, Yu RK, Avidan N, Bourgeois S, Estrera AL, Safi HJ, Sparks E, Amor D, Ades L, McConnell V, Willoughby CE, Abuelo D, Willing M, Lewis RA, Kim DH, Scherer S, Tung PP, Ahn C, Buja LM, Raman CS, Shete SS, Milewicz DM.| title =Mutations in smooth muscle alpha-actin (ACTA2) lead to thoracic aortic aneurysms and dissections| year =2007 | journal = [[Nature]] Genet. | volume =39 | number =12 | id = PMID 17994018| url =}}</ref> The number of afflictions that the gene is implicated in is increasing. It has been related to [[Moyamoya disease]] and it seems likely that certain mutations in heterozygosis could confer a predisposition to many vascular pathologies, such as thoracic aortic aneurysm and [[ischaemic heart disease]].<ref name="Guo">{{cite journal| surname =Guo | co-authors= Papke CL, Tran-Fadulu V, Regalado ES, Avidan N, Johnson RJ, Kim DH, Pannu H, Willing MC, Sparks E, Pyeritz RE, Singh MN, Dalman RL, Grotta JC, Marian AJ, Boerwinkle EA, Frazier LQ, LeMaire SA, Coselli JS, Estrera AL, Safi HJ, Veeraraghavan S, Muzny DM, Wheeler DA, Willerson JT, Yu RK, Shete SS, Scherer SE, Raman CS, Buja LM, Milewicz DM. | year=2009 |month=May | title=Mutations in smooth muscle alpha-actin (ACTA2) cause coronary artery disease, stroke, and Moyamoya disease, along with thoracic aortic disease | journal =Am J Human Genet | volume=84 | number=5| pages=617-27| pmid=19409525 }}</ref> The α-actin found in smooth muscles is also an interesting marker for evaluating the progress of liver [[cirrhosis]].<ref name="Akpolat">{{cite journal| author =Akpolat N, Yahsi S, Godekmerdan A, Yalniz M, Demirbag K| title =The value of alpha-SMA in the evaluation of hepatic fibrosis severity in hepatitis B infection and cirrhosis development: a histopathological and immunohistochemical study | year =2005 | journal = Histopathology | volume =47 | number =3 | id = PMID 16115228| url =}}</ref>

=== In heart muscle ===

The ''[[ACTC1]]'' gene codes for the α-actin isoform present in heart muscle. It was first sequenced by Hamada and co-workers in [[1982]], when it was found that it is interrupted by five introns.<ref>{{cite journal| author =Hamada H, Petrino MG, Kakunaga T.| title =Molecular structure and evolutionary origin of human cardiac muscle actin gene | year =1982 | journal = Proc Natl Acad Sci U S A. | volume =79 | number =19 | id = PMID 6310553| url =http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pubmed&pubmedid=6310553}}</ref> It was the first of the six genes where alleles were found that were implicated in pathological processes.<ref name="Olson">{{cite journal| author =Olson TM, Michels VV, Thibodeau SN, Tai YS, Keating MT.| title =Actin mutations in dilated cardiomyopathy, a heritable form of heart failure. | year =1998 | journal = [[Science]] | volume =280 | number =5364 | id = PMID 9563954}}</ref>

[[Image:Myocardiopathy dilated2.JPG|thumb|Crossection of a [[Muridae|rat]] [[heart]] that is showing signs of [[dilated cardiomyopathy]].<ref>Xia X-G, Zhou H, Samper E, Melov S, Xu Z (2006) [http://www.plosgenetics.org/article/info:doi/10.1371/journal.pgen.0020010 Pol II–Expressed shRNA Knocks Down Sod2 Gene Expression and Causes Phenotypes of the Gene Knockout in Mice. PLoS Genet 2(1): e10. doi:10.1371/journal.pgen.0020010]</ref>]]

A number of structural disorders associated with point mutations of this gene have been described that cause malfunctioning of the heart, such as Type 1R [[dilated cardiomyopathy]] and Type 11 [[hypertrophic cardiomyopathy]]. Certain defects of the [[Atrium (heart)|atrial septum]] have been described recently that could also be related to these mutations.<ref>{{cite web|url =http://www.ncbi.nlm.nih.gov/entrez/dispomim.cgi?id=102540 |title =OMIM 102540 |access date =[[27 December]]|access year =2008 }}</ref><ref name="Mattson">{{cite journal| author =Matsson H, Eason J, Bookwalter CS, Klar J, Gustavsson P, Sunnegårdh J, Enell H, Jonzon A, Vikkula M, Gutiérrez I, Granados-Riveron J, Pope M, Bu'Lock F, Cox J, Robinson TE, Song F, Brook DJ, Marston S, Trybus KM, Dahl N.| title =Alpha-cardiac actin mutations produce atrial septal defects. | year =2008 | journal = Hum Mol Genet. | volume =17 | number =2 | id = PMID 17947298 | url =}}</ref>

Two cases of dilated cardiomyopathy have been studied involving a substitution of highly conserved [[amino acid]]s belonging to the [[protein domains]] that bind and intersperse with the [[sarcomere|Z discs]]. This has led to the theory that the dilation is produced by a defect in the transmission of [[muscle contraction|contractile force]] in the [[myocyte]]s.<ref name="Olson" /><ref name=Devlin />

The mutations in”ACTC1” are responsible for at least 5% of hypertrophic cardiomyopathies.<ref>{{cite web|url =http://edoc.hu-berlin.de/dissertationen/kabaeva-zhyldyz-2002-11-11/HTML/kabaeva-ch1.html |title =Genetic analysis in hypertrophic cardiomyopathy: missense mutations in the ventricular myosin regulatory light chain gene |access date =[[27 December]]|access year =2008|author=Kabaeva, Kabaeva,Z| author link=http://edoc.hu-berlin.de/dissertationen/kabaeva-zhyldyz-2002-11-11/HTML/kabaeva-vita.html|year =2003 |month =January |format =Dissertation |publisher =[[University of Humboldt Berlin]] }}</ref> The existence of a number of point mutations have also been found:<ref>Olson TM, Doan TP, Kishimoto NY, Whitby FG, Ackerman MJ, Fananapazir L. [http://www.ncbi.nlm.nih.gov/pubmed/10966831 Inherited and de novo mutations in the cardiac actin gene cause hypertrophic cardiomyopathy]. J Mol Cell Cardiol 2000; 32:1687-1694.</ref>
* Mutation E101K: changes of net charge and formation of a weak electrostatic link in the actomyosin-binding site.
* P166A: interaction zone between actin monomers.
* A333P: actin-myosin interaction zone.

Pathogenesis appears to involve a compensatory mechanism: the mutated proteins act like “toxins” with a dominant effect, decreasing the heart’s ability to [[Heart#Functioning|contract]] causing abnormal mechanical behaviour such that the hypertrophy, that is usually delayed, is a consequence of the cardiac muscle’s normal response to [[stress]].<ref>{{cite web|url =http://www.scielo.org.ve/scielo.php?pid=S0535-51332004000100008&script=sci_arttext#tab1 |title =Cardiomiopatía hipertrófica familiar: Genes, mutaciones y modelos animales. Revisión. |access date =[[15 July]] [[2009]] | last name = Ramírez |first name =Carlos Darío|co-authors = Raúl Padrón|year =2004 |month =March|publisher = Invest. clín, mar. 2004, vol.45, no.1, p.69-100 |language =Spanish}}</ref>

Recent studies have discovered “ACTC1” mutations that are implicated in two other pathological processes: Infantile idiopathic [[restrictive cardiomyopathy]],<ref name="RCM">{{cite journal| author =Kaski JP, Syrris P, Burch M, Tomé-Esteban MT, Fenton M, Christiansen M, Andersen PS, Sebire N, Ashworth M, Deanfield JE, McKenna WJ, Elliott PM.| title =Idiopathic restrictive cardiomyopathy in children is caused by mutations in cardiac sarcomere protein genes | year =2008 | journal = Heart | volume =94 | number =11 | id = PMID 18467357 | url =}}</ref> and [[Noncompaction cardiomyopathy|noncompaction of the left ventricular myocardium]].<ref name="nocompacto">{{cite journal| author =Klaassen S, Probst S, Oechslin E, Gerull B, Krings G, Schuler P, Greutmann M, Hürlimann D, Yegitbasi M, Pons L, Gramlich M, Drenckhahn JD, Heuser A, Berger F, Jenni R, Thierfelder L.| title =Mutations in sarcomere protein genes in left ventricular noncompaction | year =2008 | journal = Circulation | volume =117 | number =22 | id = PMID 1850600}}</ref>

=== In cytoplasmatic actins ===
''[[ACTB]]'' is a highly complex [[Locus (genetics)|locus]]. A number of [[pseudogenes]] exist that are distributed throughout the [[genome]], and its sequence contains six exons that can give rise to up to 21 different transcriptions by [[alternative splicing]], which are known as the β-actins. Consistent with this complexity, its products are also found in a number of locations and they form part of a wide variety of processes ([[cytoskeleton]], NuA4 [[histone]]-acyltransferase complex, [[cell nucleus]]) and in addition they are associated with the mechanisms of a great number of pathological processes ([[Cancer|carcinomas]], juvenile [[dystonia]], infection mechanisms, [[nervous system]] malformations and tumour invasion, among others).<ref>{{cite web| url =http://www.ncbi.nlm.nih.gov/IEB/Research/Acembly/av.cgi?db=human&l=ACTB |title =ACEVieW:Homo sapiens complex locus ACTB, encoding actin, β. |access date =28 June |access year =2008 }}</ref> A new form of actin has been discovered, kappa actin, which appears to substitute for β-actin in processes relating to [[tumour]]s.<ref>{{cite journal| author =Chang KW, Yang PY, Lai HY, Yeh TS, Chen TC, Yeh CT.| title =Identification of a novel actin isoform in hepatocellular carcinoma | year =2006 | journal = Hepatol Res. | volume =36 | number =1 | id = PMID 16824795}}</ref>

[[Image:Neuron actin cytoskeleton.JPG|left|thumb|Image taken using [[confocal microscopy]] and employing the use of specific [[Antibody|antibodies]] showing actin’s cortical network. In the same way that in juvenile [[dystonia]] there is an interruption in the structures of the [[cytoskeleton]], in this case it is produced by [[cytochalasin D]].<ref>Kristy L Williams, Masuma Rahimtula and Karen M Mearow. [http://www.biomedcentral.com/1471-2202/6/24 Hsp27 and axonal growth in adult sensory neurons in vitro] BMC Neuroscience 2005, 6:24doi:10.1186/1471-2202-6-24</ref>]]

Three pathological processes have so far been discovered that are caused by a direct alteration in gene sequence:

* [[Hemangiopericytoma]] with t(7;12)(p22;q13)-translocations is a rare affliction, in which a [[Mutation#By effect on structure|translocational mutation]] causes the fusion of the ''ACTB'' gene over [[GLI1]] in [[Chromosome 12 (human)|Chromosome 12]].<ref>{{cite web|url =http://atlasgeneticsoncology.org/Tumors/Pericytomt0712ID5192.html |title =Pericytoma with t(7;12) |access date =[[28 December]]|access year =2008|publisher =Atlas of Genetics and Cytogenetics in Oncology and Haematology}}</ref>

* Juvenile onset [[dystonia]] is a rare [[degenerative disease]] that affects the [[central nervous system]], in particular its affects areas of the [[neocortex]] and [[thalamus]], where rod-like [[eosina|eosinophilic]] inclusions are formed. The affected individuals represent a [[phenotype]] with deformities on the median line, sensory [[Deafness|hearing loss]] and dystonia. It is caused by a point mutation in which the amino acid [[tryptophan]] replaces [[arginine]] in position 183. This alters actin’s interaction with the ADF/[[cofilin]] system, which regulates the dynamics of [[Neuron|nerve cell]] cytoskeleton formation.<ref name="procaccio">{{cite journal| author =Procaccio V, Salazar G, Ono S, Styers ML, Gearing M, Dávila A, Jiménez R, Juncos J, Gutekunst CA, Meroni G, Fontanella B, Sontag E, Sontag JM, Faundez V, Wainer BH.| title =A mutation of β -actin that alters depolymerization dynamics is associated with autosomal dominant developmental malformations, deafness, and dystonia | year =2006 | journal = Am J Hum Genet | volume =78 | number =6 | id = PMID 16685646| url =http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pubmed&pubmedid=16685646#RF21}}</ref>

* A dominant point mutation has also been discovered that causes [[neutrophil granulocyte]] dysfunction and recurring [[infection]]s. It appears that the mutation modifies the domain responsible for binding between [[profilin]] and other regulatory proteins. Actin’s affinity for profilin is greatly reduced in this allele.<ref>{{cite journal| author =Nunoi H, Yamazaki T, Tsuchiya H, Kato S, Malech HL, Matsuda I, Kanegasaki S.| title =A heterozygous mutation of β-actin associated with neutrophil dysfunction and recurrent infection | year =1999 | journal = Proc Natl Acad Sci U S A. | volume =96 | number =15 | id = PMID 10411937| url =http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pubmed&pubmedid=10411937}}</ref>

The ''[[ACTG1]]'' locus codes for the cytosolic γ-actin protein that is responsible for the formation of cytoskeletal [[microfilament]]s. It contains 6 [[exon]]s, giving rise to 22 different [[Messenger RNA|mRNAs]], which produce 4 complete [[isoform]]s whose form of expression is probably dependent on the type of [[Tissue (biology)|tissue]] they are found in. It also has two different [[Promoter (genetics)|DNA promoter]]s. <ref name="actgview">{{cite web|url =http://www.ncbi.nlm.nih.gov/IEB/Research/Acembly/av.cgi?db=human&l=ACTG1 |title =ACEVIEW ACTG1 |access date =[[29 December]]|access year =2008 }}</ref> It has been noted that the sequences translated from this locus and from that of β-actin are very similar to the predicted ones, suggesting a common ancestral sequence that suffered duplication and genetic conversion.<ref>{{cite journal| author =Erba HP, Gunning P, Kedes L.| title =Nucleotide sequence of the human γ-cytoskeletal actin mRNA: anomalous evolution of vertebrate non-muscle actin genes. | year =1986 | journal = Nucleic Acids Res. | volume =14 | number =13 | id = PMID 373740| url =http://www.pubmedcentral.nih.gov/pagerender.fcgi?artid=311540&pageindex=2#page}}</ref>

In terms of pathology, it has been associated with processes such as [[amyloidosis]], [[retinitis pigmentosa]], infection mechanisms, [[kidney]] diseases and various types of congenital hearing loss.<ref name="actgview" />

Six autosomal-dominant point mutations in the sequence have been found to cause various types of hearing loss, particularly sensorineural hearing loss linked to the DFNA 20/26 locus. It seems that they affect the [[stereocilia]] of the ciliated cells present in the inner ear’s [[Organ of Corti]]. β-actin is the most abundant protein found in human tissue, but it is not very abundant in ciliated cells, which explains the location of the pathology. On the other hand, it appears that the majority of these mutations affect the areas involved in linking with other proteins, particularly actomyosin.<ref name="dosremedios" /> Some experiments have suggested that the pathological mechanism for this type of hearing loss relates to that fact that the F-actin in the mutations is more sensitive to cofilin than normal.<ref name="Bryan">{{cite journal| surname =Bryan | name =KE |co-authors=Rubenstein, PA | year=2009 |month= July | title =Allele-specific Effects of Human Deafness {γ}-Actin Mutations (DFNA20/26) on the Actin/Cofilin Interaction | journal =J Biol Chem. | volume=284 | number=27 | pages=18260-9 |pmid=19419963}}</ref>

However, although there is no record of any case, it is know that γ-actin is also expressed in skeletal muscles, and although it is present in small quantities, [[model organism]]s have shown that its absence can give rise to [[myopathy]]s.<ref>{{cite journal| author =Sonnemann KJ, Fitzsimons DP, Patel JR, Liu Y, Schneider MF, Moss RL, Ervasti JM.
| title =Cytoplasmic γ-actin is not required for skeletal muscle development but its absence leads to a progressive myopathy. | year =2006 | journal = Dev Cell | volume =11 | number =3 | id = PMID 16950128| url =}}</ref>

=== Other pathological mechanisms ===

Some infectious agents use actin, especially cytoplasmic actin, in their [[Alternation of generations|life cycle]]. Two basic forms are present in [[bacteria]]:
* ''[[Listeria monocytogenes]]'', some species of ''[[Rickettsia]]'', ''[[Shigella flexneri]]'' and other intracellular germs escape from [[phagocytosis|phagocytic]] vacuoles by coating themselves with a capsule of actin filaments. ''L. monocytogenes'' and ''S. flexneri'' both generate a tail in the form of a "comet tail" that gives them mobility. Each species exhibits small differences in the molecular polymerization mechanism of their «comet tails». Different displacement velocities have been observed, for example, with ''Listeria'' and ''Shigella'' found to be the fastest.<ref name=Gouin1999>{{Cite journal | surname= Gouin | name= E. | year= 1999 |title = A comparative study of the actin-based motilities of the pathogenic bacteria Listeria monocytogenes, Shigella flexneri and Rickettsia conorii | pub-journal= Journal of Cell Science | volume= 112 | number= 11 | page = 1697–1708 | url = http://jcs.biologists.org/cgi/reprint/112/11/1697.pdf | format= w}}</ref> Many experiments have demonstrated this mechanism ''in vitro''. This indicates that the bacteria are not using a myosin-like protein motor, and it appears that their propulsion is acquired from the pressure exerted by the polymerization that takes place near to the microorganism's cell wall. The bacteria have previously been surrounded by ABPs from the host, and as a minimum the covering contains [[Arp2/3 complex]], [[Ena/Vasp homology proteins|Ena/VASP proteins]], cofilin, a buffering protein and nucleation promoters, such as [[vinculin]] complex. Through these movements they form protrusions that reach the neighbouring cells, infecting them as well so that the [[immune system]] can only fight the infection through cell immunity. The movement could be caused by the modification of the curve and debranching of the filaments.<ref name="Lambrechts">{{cite journal| surname =Lambrechts | name =A| co-authors= Gevaert K, Cossart P, Vandekerckhove J, Van Troys M. | year=2008 |month=May | title =Listeria comet tails: the actin-based motility machinery at work | journal =Trends Cell Biol. | volume=18 | number=5 | pages=220-7| pmid=18396046 }}</ref> Other species, such as ''[[Mycobacterium marinum]]'' and ''[[Burkholderia pseudomallei]]'', are also capable of localized polymerization of cellular actin to aid their movement through a mechanism that is centered on the Arp2/3 complex. In addition the vaccine [[virus]] ''[[Vaccinia]]'' also uses elements of the actin cytoskeleton for its dissemination.<ref name=Gouin2005>{{Cite journal | surname1= Gouin | name1= E. | surname2= Welch | name2= M.D. | surname3= Cossart | name3= P. | year= 2005 | title = Actin-based motility of intracellular pathogens | pub-journal= Current Opinion in Microbiology | volume= 8 | number= 1 | page = 35–45 | url = http://linkinghub.elsevier.com/retrieve/pii/S1369527404001675}}</ref>
* ''[[Pseudomonas aeruginosa]]'' is able to form a protective [[biofilm]] in order to escape a [[host (biology)|host organism]]’s defences, especially [[Neutrophil granulocyte|white blood cells]] and [[Antibacterial|antibiotics]]. The biofilm is constructed using [[DNA]] and actin filaments from the host organism.<ref name="Parks">{{cite journal| surname =Parks | name =QM| co-authors=Young RL, Poch KR, Malcolm KC, Vasil ML, Nick JA.| year=2009 |month= April | title =Neutrophil enhancement of Pseudomonas aeruginosa biofilm development: human F-actin and DNA as targets for therapy | journal =J med microbiol | pages=492-502 pmid=19273646 | doi=10.1099/jmm.0.005728-0 |url=http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pubmed&pubmedid=19273646}}</ref>


In addition to the previously cited example, actin polymerization is stimulated in the initial steps of the internalization of some viruses, notably [[HIV]], by, for example, inactivating the cofilin complex.<ref name="LiuY">{{cite journal| surname =Liu | co-authors= Belkina NV, Shaw S. | year=2009 |month= April | title =HIV infection of T cells: actin-in and actin-out | journal =Sci Signal. | volume=2 | pmid=19366992 }}</ref>

The role that actin plays in the invasion process of cancer cells has still not been determined.<ref name="Machesky">{{cite journal| surname =Machesky LM | co-authors=Tang HR || year=2009 |month= July | title =Actin-Based Protrusions: Promoters or Inhibitors of Cancer Invasion? | journal =Cancer Cell. | volume=16 | number=1 | pages=5-7 || pmid= 19573806}}</ref>

== Evolution ==
The eukaryotic cytoskeleton of organisms among all [[Phylogenetic|taxonomic groups]] have similar components to actin and tubulin. For example, the protein that is coded by the ''[[ACTG2]]'' gene in humans is completely equivalent to the [[Homology (biology)|homologues]] present in rats and mice, even though at a [[nucleotide]] level the similarity decreases to 92 %. <ref name=Miwa1991>{{Cite journal | surname= Miwa | name= T. Manabe | year= 1991 | title = The nucleotide sequence of a human smooth muscle (enteric type) γ-actin cDNA. | pub-journal= Molecular and Cellular Biology | volume= 11 | number= 6 | page = 3296–3306 | url = http://www.pubmedcentral.nih.gov/picrender.fcgi?artid=360182}}</ref> However, there are major differences with the equivalents in prokaryotes ([[FtsZ]] and [[MreB]]), where the similarity between nucleotide sequences is between 40−50 % among different [[bacteria]] and [[archaea]] species. Some authors suggest that the ancestral protein that gave rise to the model eukaryotic actin resembles the proteins present in modern bacterial cytoskeletons.<ref name=Erickson2007>{{Cite journal| surname= Erickson | name= H.P| year= 2007| title = Evolution of the cytoskeleton| pub-journal= BioEssays: news and reviews in molecular, cellular and developmental biology| volume= 29| number= 7| page = 668| url = http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2630885}}</ref>

[[Image:MreB.png|thumb|Structure of [[MreB]], a bacterial protein whose three-dimensional structure resembles that of G-actin.]]

Some authors point out that the behaviour of actin, tubulin and [[histone]], a protein involved in the stabilization and regulation of DNA, are similar in their ability to bind nucleotides and in their ability of take advantage of [[Brownian motion]]. It has also been suggested that they all have a common ancestor.<ref name=Gardiner2008>{{Cite journal| surname= Gardiner | name= J| year= 2008 | title = Are histones, tubulin, and actin derived from a common ancestral protein?| pub-journal= Protoplasma| volume= 233| page = 1 | doi = 10.1007/s00709-008-0305-z}}</ref> Therefore [[Evolution|evolutionary]] processes resulted in the diversification of ancestral proteins into the varieties present today, conserving, among others, actins as efficient molecules that were able to tackle essential ancestral biological processes, such as [[endocytosis]].<ref>{{Cite journal| surname= J.A| year= 2009| title = Actin and endocytosis: mechanisms and phylogeny| pub-journal= Current Opinion in Cell Biology| volume= 21| page = 20| doi = 10.1016/j.ceb.2009.01.0 | url = http://linkinghub.elsevier.com/retrieve/pii/S0955067409000076}}</ref>

=== Equivalents in bacteria ===
The [[Cytoskeleton#The prokaryotic cytoskeleton|bacterial cytoskeleton]] may not be as complex as that found in [[eukaryote]]s, however, it contains proteins that are highly similar to actin monomers and polymers. The bacterial protein [[MreB]] polymerizes into thin non-helical filaments and occasionally into helical structures similar to F-actin. <ref name="Toshiro" /> Furthermore its crystalline structure is very similar to that of G-actin (in terms of its three dimensional conformation), there are even similarities between the MreB protofilaments and F-actin. The bacterial cytoskeleton also contains the [[FtsZ]] proteins, which are similar to [[tubulin]].<ref name=Van2001>{{Cite journal | surname1= Van Den Ent | name1= F. | surname2= Amos | name2= L.A. | surname3= Löwe | name3= J. | year= 2001 | title = Prokaryotic origin of the actin cytoskeleton | pub-journal= Nature | volume= 413 | page = 39–44 | doi = 10.1038/35092500 | url = http://www2.mrc-lmb.cam.ac.uk/ss/Amos_L/group/PDF/MreB%20Nature.pdf | format= w}}</ref>

Bacteria therefore possess a cytoskeleton with homologous elements to actin (for example, MreB, ParM, and MamK), even though the amino acid sequence of these proteins diverges from that present in animal cells. However, MreB and ParM have a high degree of [[Protein structure|structural]] similarity to eukaryotic actin. The highly dynamic microfilaments formed by the aggregation of MreB and ParM are essential to cell viability and they are involved in cell morphogenesis, [[chromosome]] segregation and cell polarity. ParM is an actin homologue that is coded in a [[plasmid]] and it is involved in the regulation of plasmid DNA.<ref name=Carballido-lopez2006>{{Cite journal | surname= Carballido-lopez | name= Rut | year= 2006 | title = The Bacterial Actin-Like Cytoskeleton | pub-journal= Microbiology and Molecular Biology Reviews | volume= 70 | number= 4 | page = 888–909 | doi = 10.1128/MMBR.00014-06 | url = http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=1698507 | pmid = 17158703 | format= w}}</ref>


==History==
== Applications ==
Actin is used in scientific and technological laboratories as a track for [[molecular motor]]s such as myosin (either in muscle tissue or outside it) and as a necessary component for cellular functioning. It can also be used as a diagnostic tool, as several of its anomalous variants are related to the appearance of specific pathologies.
Actin was first observed [[experiment]]ally in 1887 by [[W.D. Halliburton]], who extracted a protein from muscle that 'coagulated' preparations of myosin, which he dubbed "myosin-ferment".<ref name="Halliburton">{{cite journal |author=Halliburton, W.D. |title=On muscle plasma |journal=J. Physiol. |volume=8 |issue= 3-4|pages=133 |year=1887 |pmid=16991477 |pmc=1485127 }}</ref> However, Halliburton was unable to further characterise his findings, and the discovery of actin is credited instead to [[Brunó Ferenc Straub]], a young biochemist working in [[Albert Szent-Györgyi]]'s laboratory at the Institute of Medical Chemistry at the [[University of Szeged]], [[Hungary]].


* [[Nanotechnology]]. Actin-myosin systems act as molecular motors that permit the transport of vesicles and organelles throughout the cytoplasm. It is possible that actin could be applied to [[nanotechnology]] as its dynamic ability has been harnessed in a number of experiments including those carried out in acellular systems. The underlying idea is to use the microfilaments as tracks to guide molecular motors that can transport a given load. That is actin could be used to define a circuit along which a load can be transported in a more or less controlled and directed manner. In terms of general applications, it could be used for the directed transport of molecules for deposit in determined locations, which would permit the controlled assembly of nanostructures.<ref>{{Cite journal| surname= J| year= 2001| title = Light-controlled molecular shuttles made from motor proteins carrying cargo on engineered surfaces| pub-journal= Nano Letters| volume= 1| number= 5| page = 235–239| doi = 10.1021/nl015521e}}</ref> These attributes could be applied to laboratory processes such as on ''[[lab-on-a-chip]]'', in nanocomponent mechanics and in nanotransformers that convert mechanical energy into electrical energy.<ref name=Mansson2005>{{Cite journal| surname= Mansson | name= A. Sundberg| year= 2005| title = Actin-Based Molecular Motors for Cargo Transportation in Nanotechnology—Potentials and Challenges| pub-journal= IEEE Transactions on Advanced Packaging| volume= 28| number= 4| page = 547–555| url = http://ieeexplore.ieee.org/xpls/abs_all.jsp?arnumber=1528636}}</ref>
In 1942, Straub developed a novel technique for extracting muscle protein that allowed him to isolate substantial amounts of relatively pure actin. Straub's method is essentially the same as that used in laboratories today. Szent-Gyorgyi had previously described the more viscous form of myosin produced by slow muscle extractions as 'activated' myosin, and, since Straub's protein produced the activating effect, it was dubbed ''actin''. The hostilities of [[World War II]] meant Szent-Gyorgyi and Straub were unable to publish the work in [[Western countries|Western]] [[scientific journal]]s; it became well known in the West only in 1945, when it was published as a supplement to the ''Acta Physiologica Scandinavica''.<ref name="Szent_Gyorgyi">{{cite journal |author=Szent-Gyorgyi, A. |title=Studies on muscle |journal=Acta Physiol Scandinav |volume=9 |issue=Suppl |pages=25 |year=1945 }}</ref>


* Internal control of techniques used in [[molecular biology]], such as ''[[western blot]]'' and [[real-time polymerase chain reaction]]. As actin is essential for cell survival it has been postulated that the quantity of actin is under such tight control at a cellular level that it can be assumed that its [[Transcription (genetics)|transcription]] (that is, the degree to which its [[gene]]s are expressed) and [[Translation (biology)|translation]], that is the production of protein, is practically constant and independent of experimental conditions. Therefore, it is common practice in protein quantification studies (''western blot'') and [[Messenger RNA|transcription]] studies (Real-time polymerase chain reaction) to carry out the quantification of the gene of interest and also the quantification of a reference gene such as the one that codes for actin. By dividing the quantity of the gene of interest by that of the actin gene it is possible to obtain a relative quantity that can be compared between different experiments,<ref>Vandesompele J, De Preter K, Pattyn F, Poppe B, Van Roy N, De Paepe A, Speleman F (2002) [http://genomebiology.com/2002/3/7/RESEARCH/0034./ABSTRACT/ADDITIONAL/ABSTRACT/COMMENTS/additional/ Accurate normalisation of real-time quantitative RT-PCR data by geometric averaging of multiple internal control genes]. Gen. Biol. 3: 1–12.</ref> whenever the expression of the latter is constant. It is worth pointing out that actin does not always have the desired stability in its [[gene expression]].<ref>{{Cite journal | title = ß-Actin | url = http://linkinghub.elsevier.com/retrieve/pii/S0890850801903767 | year= 2001 | pub-journal= Molecular and cellular Probes | page = 307–311 | volume= 15 | number= 5 | surname1= Selvey | name1= S. | surname2= Thompson| name2= E.W. | surname3= Matthaei | name3= K. | surname4= Lea | name4= R.A. | surname5= Irving| name5= M.G. | surname6= Griffiths| name6= L.R. | access date= 7 July 2009}}</ref>
Straub continued to work on actin, and in 1950 reported that actin contains bound [[adenosine triphosphate|ATP]] <ref name="Straub">{{cite journal |author=Straub FB, Feuer G |title=Adenosinetriphosphate. The functional group of actin. 1950 |journal=Biochim. Biophys. Acta |volume=1000 |issue= |pages=180–95 |year=1989 |pmid=2673365 }}</ref> and that, during polymerisation of the protein into microfilaments, the [[nucleotide]] is hydrolysed to [[adenosine diphosphate|ADP]] and inorganic [[phosphate]] (which remain bound in the microfilament). Straub suggested the transformation of ATP-bound actin to ADP-bound actin played a role in muscular contraction. In fact, this is true only in [[smooth muscle]], and was not supported through experimentation until 2001.<ref name="Bárány">{{cite journal |author=Bárány M, Barron JT, Gu L, Bárány K |title=Exchange of the actin-bound nucleotide in intact arterial smooth muscle |journal=J. Biol. Chem. |volume=276 |issue=51 |pages=48398–403 |year=2001 |month=December |pmid=11602582 |doi=10.1074/jbc.M106227200 |url=http://www.jbc.org/cgi/pmidlookup?view=long&pmid=11602582}}</ref>


* Health. Some [[allele]]s of actin cause diseases, for this reason techniques for their detection have been developed. In addition, actin can be used as an indirect marker in surgical pathology: it is possible to use variations in the pattern of its distribution in tissue as a marker of invasion in [[Neoplasm|neoplasia]], [[vasculitis]] and other conditions.<ref>{{Cite journal| title = Immunohistochemical localization of actin: Applications in surgical pathology | url = http://www.ajsp.com/pt/re/ajsp/abstract.00000478-198101000-00013.htm | title = Immunohistochemical localization of actin: Applications in surgical pathology. | year= 1981| pub-journal= The American Journal of Surgical Pathology | page = 91 | volume= 5 | number= 1| surname1= Mukai | name1= K. | surname2= Schollmeyer| name2= J.V. | surname3= Rosai| name3= J. | access date= 2009-04-19}}</ref> Further, due to actin’s close association with the apparatus of muscular contraction its levels in skeletal muscle diminishes when these tissues [[atrophy]], it can therefore be used as a marker of this physiological process. <ref>{{Cite journal| title = Atrophy responses to muscle inactivity. II. Molecular markers of protein deficits| url = http://jap.physiology.org/cgi/content/full/95/2/791| title = Atrophy responses to muscle inactivity. II. Molecular markers of protein deficits | year= 2003| pub-journal= Journal of Applied Physiology | page = 791–802 | volume= 95 | number= 2 | surname1= Haddad | name1= F. | surname2= Roy| name2= R.R. | surname3= Zhong| name3= H. | surname4= Edgerton| name4= V.R. | surname5= Baldwin | name5= K.M. | access date= 19 March 2009}}</ref>
The [[X-ray crystallography|crystal structure]] of G-actin was solved in 1990 by Kabsch and colleagues.<ref name="Kabsch">{{cite journal |author=Kabsch W, Mannherz HG, Suck D, Pai EF, Holmes KC |title=Atomic structure of the actin:DNase I complex |journal=Nature |volume=347 |issue=6288 |pages=37–44 |year=1990 |month=September |pmid=2395459 |doi=10.1038/347037a0}}</ref> In the same year, a model for F-actin was proposed by Holmes and colleagues.<ref name="Holmes">{{cite journal |author=Holmes KC, Popp D, Gebhard W, Kabsch W |title=Atomic model of the actin filament |journal=Nature |volume=347 |issue=6288 |pages=44–9 |year=1990 |month=September |pmid=2395461 |doi=10.1038/347044a0}}</ref> The model was derived by fitting a helix of G-actin structures according to low-resolution fiber diffraction data from the filament. Several models of the filament have been proposed since. However, there is still no high-resolution X-ray structure of F-actin.


* [[Food technology]]. It is possible to determine the quality of certain processed foods, such as [[Embutido|sausages]], by quantifying the amount of actin present in the constituent [[meat]]. Traditionally, a method has been used that is based on the detection of [[Histidine|3-methylhistidine]] in [[hydrolysis|hydrolyzed]] samples of these products, as this compound is present in actin and F-myosin’s heavy chain (both are major components of muscle). The generation of this compound in animal flesh derives from the [[methylation]] of [[histidine]] residues present in both proteins.<ref>{{Cite journal | title = Current advances in proteomic analysis and its use for the resolution of poultry meat quality | url = http://www.animalscience.com/uploads/additionalFiles/QualityOfPoultryMeat/27.pdf | year= 2006 | pub-journal= World's Poultry Science Journal | page = 123–130 | volume= 62 | number= 01 | surname1= Remignon | name1= H. | surname2= Molette| name2= C. | surname3= Babile| name3= R. | surname4= Fernandez| name4= X. | access date= 7 July 2009}}</ref><ref>{{Cite journal | title = Methods and Instruments in Applied Food Analysis | url = http://orton.catie.ac.cr/cgi-bin/wxis.exe/?IsisScript=BFHIA.xis | year= 2004 | author= M.L. | pub-journal= Handbook of Food Analysis. | access date= 7 July 2009}}</ref>
The ''[[Listeria]]'' bacteria use the cellular machinery to move around inside the host cell, by inducing directed polymerisation of actin by the [[Actin assembly-inducing protein|ActA]] [[transmembrane protein]], thus pushing the bacterial cell around.


==See also==
== See also ==
* [[Actin-binding protein]]
* [[Arp2/3]]
* [[Phallotoxin]]
* [[Filopodia]]
* [[Intermediate filament]]
* [[FtsZ]]
* [[Lamellipodium]]
*[[MreB]] — one of the actin homologues in bacteria
*[[MreB]] — one of the actin homologues in bacteria
*[[Motor protein]] — converts chemical energy into mechanical work
*[[Motor protein]] — converts chemical energy into mechanical work
Line 84: Line 359:
*[[Active matter]]
*[[Active matter]]


==References==
== Notes ==
{{listaref|group=n.}}
<!--See [[Wikipedia:Footnotes]] for an explanation of how to generate footnotes using the <ref(erences/)> tags-->

{{reflist}}
== Referencias ==
{{listaref|2}}


{{Cytoskeletal proteins}}
{{Cytoskeletal proteins}}
{{Muscle tissue}}
{{Muscle tissue}}
{{Autoantigens}}
{{Autoantigens}}

==External links==
* [http://www.mechanobio.info/Home/glossary-of-terms/mechano-glossary--a/mechano-glossary-actin MBInfo - Actin]
* [http://www.mechanobio.info/Home/essential-info/What-is-the-Role-of-Actin-Filaments-in-Mechanotransduction MBInfo - Actin in Mechanobiology]
*[http://www.pdbe.org/emsearch/actin* 3D macromolecular structures of actin filaments from the EM Data Bank(EMDB)]


[[Category:Cytoskeleton]]
[[Category:Cytoskeleton]]
[[Category:Structural proteins]]
[[Category:Structural proteins]]
[[Category:Autoantigens]]
[[Category:Autoantigens]]
[[Category:Article Feedback 5]]


{{Link FA|es}}
{{Link FA|es}}
==External links==
* [http://www.mechanobio.info/Home/glossary-of-terms/mechano-glossary--a/mechano-glossary-actin MBInfo - Actin]
* [http://www.mechanobio.info/Home/essential-info/What-is-the-Role-of-Actin-Filaments-in-Mechanotransduction MBInfo - Actin in Mechanobiology]
*[http://www.pdbe.org/emsearch/actin* 3D macromolecular structures of actin filaments from the EM Data Bank(EMDB)]


[[ar:أكتين]]
[[ar:أكتين]]
Line 109: Line 386:
[[da:Actin]]
[[da:Actin]]
[[de:Aktin]]
[[de:Aktin]]
[[et:Aktiin]]
[[es:Actina]]
[[es:Actina]]
[[et:Aktiin]]
[[eu:Aktina]]
[[eu:Aktina]]
[[fa:اکتین]]
[[fa:اکتین]]
[[fi:Aktiini]]
[[fr:Actine]]
[[fr:Actine]]
[[ga:Aictin]]
[[ga:Aictin]]
[[gl:Actina]]
[[gl:Actina]]
[[he:אקטין]]
[[ht:Aktin]]
[[id:Aktin]]
[[id:Aktin]]
[[it:Actina]]
[[it:Actina]]
[[he:אקטין]]
[[ja:アクチン]]
[[jv:Aktin]]
[[kk:Актин]]
[[kk:Актин]]
[[ht:Aktin]]
[[mk:Актиниум]]
[[mk:Актиниум]]
[[nl:Actine]]
[[nl:Actine]]
[[ja:アクチン]]
[[pl:Aktyna]]
[[pl:Aktyna]]
[[pt:Actina]]
[[pt:Actina]]
[[ro:Actină]]
[[ro:Actină]]
[[ru:Актин]]
[[ru:Актин]]
[[sh:Aktinski filamenti]]
[[sk:Aktín]]
[[sq:Aktini]]
[[sq:Aktini]]
[[sk:Aktín]]
[[sr:Актински филаменти]]
[[sr:Актински филаменти]]
[[sh:Aktinski filamenti]]
[[fi:Aktiini]]
[[sv:Aktin]]
[[sv:Aktin]]
[[th:แอกติน]]
[[tl:Actin]]
[[tl:Actin]]
[[th:แอกติน]]
[[tr:Aktin]]
[[tr:Aktin]]
[[uk:Актин]]
[[uk:Актин]]

Revision as of 10:25, 15 January 2013

G-actin (PDB code: 1j6z). ADP and the divalent cation are highlighted.
Actin
Identifiers
SymbolActin
PfamPF00022
InterProIPR004000
PROSITEPDOC00340
SCOP22btf / SCOPe / SUPFAM
Available protein structures:
Pfam  structures / ECOD  
PDBRCSB PDB; PDBe; PDBj
PDBsumstructure summary
PDBPDB: 1atn​ actin from Oryctolagus cuniculus


F-actin; surface representation of a repetition of 13 subunits based on Ken Holmes' actin filament model.[1]

Actin is a globular multi-functional protein that forms microfilaments. It is found in all eukaryotic cells (the only known exception being nematode sperm), where it may be present at concentrations of over 100 μM. Actin is roughly 42-kDa in size and it is the monomeric subunit of two types of filaments in cells: microfilaments, one of the three major components of the cytoskeleton, and thin filaments, part of the contractile apparatus in muscle cells. It can be present as either a free monomer called G-actin or as part of a linear polymer microfilament called F-actin both of which are essential for such important cellular functions as the mobility and contraction of cells during cell division.

Thus, actin participates in many important cellular processes, including muscle contraction, cell motility, cell division and cytokinesis, vesicle and organelle movement, cell signalling, and the establishment and maintenance of cell junctions and cell shape. Many of these processes are mediated by extensive and intimate interactions of actin with cellular membranes.[2] In vertebrates, three main groups of actin isoforms, alpha, beta, and gamma have been identified. The alpha actins, found in muscle tissues, are a major constituent of the contractile apparatus. The beta and gamma actins coexist in most cell types as components of the cytoskeleton, and as mediators of internal cell motility.

Its amino acid sequence is also one of the most highly-conserved of the proteins as it has changed little over the course of evolution, differing by no more than 20% in species as diverse as algae and humans. It is therefore considered to have an optimised structure. It has two distinguishing features: it is an enzyme that slowly hydrolizes ATP, the "universal energy currency" of biological processes. However, the ATP is required in order to maintain its structural integrity. Its efficient structure is formed by an almost unique folding process. In addition, it is able to carry out more interactions than any other protein, which allows it to perform a wider variety of functions than other proteins at almost every level of cellular life. Myosin is an example of a protein that bonds with actin. Another example is villin, which can weave actin into bundles or cut the filaments depending on the concentration of calcium cations in the surrounding medium.[3]

A cell’s ability to dynamically form microfilaments provides the scaffolding that allows it to rapidly remodel itself in response to it’s environment or to the organism’s internal signals, for example, to increase cell membrane absorption or increase cell adhesion in order to form cell tissue. Other enzymes or organelles such as cillia can be anchored to this scaffolding in order to control the deformation of the external cell membrane, which allows endocytosis and cytokinesis. It can also produce movement either by itself or with the help of molecular motors. Actin therefore contributes to processes such as the intracellular transport of vesicles and organelles as well as muscular contraction and cellular migration. It therefore plays an important role in embryogenesis, the healing of wounds and the invasivity of cancer cells. The evolutionary origin of actin can be traced to prokaryotic cells, which have equivalent proteins. Lastly, actin plays an important role in the control of gene expression.

A large number of illnesses and diseases are caused by mutations in alleles of the genes that regulate the production of actin or of its associated proteins. The production of actin is also key to the process of infection by some pathogenic microorganisms. Mutations in the different genes that regulate actin production in humans can cause muscular diseases, variations in the size and function of the heart as well as deafness. The make-up of the cytoskeleton is also related to the pathogenicity of intracellular bacteria and viruses, particularly in the processes related to evading the actions of the immune system.[4]

History

Nobel Prize winning physiologist Albert von Szent-Györgyi Nagyrápolt, co-discoverer of actin with Brunó Ferenc Straub.

Actin was first observed experimentally in 1887 by W.D. Halliburton, who extracted a protein from muscle that 'coagulated' preparations of myosin that he called "myosin-ferment".[5] However, Halliburton was unable to further refine his findings, and the discovery of actin is credited instead to Brunó Ferenc Straub, a young biochemist working in Albert Szent-Györgyi's laboratory at the Institute of Medical Chemistry at the University of Szeged, Hungary.

In 1942, Straub developed a novel technique for extracting muscle protein that allowed him to isolate substantial amounts of relatively pure actin. Straub's method is essentially the same as that used in laboratories today. Szent-Gyorgyi had previously described the more viscous form of myosin produced by slow muscle extractions as 'activated' myosin, and, since Straub's protein produced the activating effect, it was dubbed actin. Adding ATP to a mixture of both proteins (called actomyosin) causes a decrease in viscosity. The hostilities of World War II meant Szent-Gyorgyi and Straub were unable to publish the work in Western scientific journals. Actin therefore only became well known in the West in 1945, when their paper was published as a supplement to the Acta Physiologica Scandinavica.[6] Straub continued to work on actin, and in 1950 reported that actin contains bound ATP [7] and that, during polymerization of the protein into microfilaments, the nucleotide is hydrolyzed to ADP and inorganic phosphate (which remain bound to the microfilament). Straub suggested that the transformation of ATP-bound actin to ADP-bound actin played a role in muscular contraction. In fact, this is true only in smooth muscle, and was not supported through experimentation until 2001.[7][8]

The amino acid sequencing of actin was completed by M. Elzinga and co-workers in 1973,[9]. The crystal structure of G-actin was solved in 1990 by Kabsch and colleagues.[10] In the same year, a model for F-actin was proposed by Holmes and colleagues following experiments using co-crystallization with different proteins.[11] The procedure of co-crystallization with different proteins was used repeatedly during the following years, until in 2001 the isolated protein was crystallized along with ADP. However, there is still no high-resolution X-ray structure of F-actin. The crystallization of F-actin was possible due to the use of a rhodamine conjugate that impedes polymerization by blocking the amino acid cys-374.[12] Christine Oriol-Audit died in the same year that actin was first crystalized but she was the researcher that in 1977 first crystalized actin in the absence of Actin Binding Proteins (ABPs). However, the resulting crystals were too small for the available technology of the time.[13]

Although no high-resolution model of actin’s filamentous form currently exists, in 2008 Sawaya’s team were able to produce a more exact model of its structure based on multiple crystals of actin dimers that bind in different places.[14] This model has subsequently been further refined by Sawaya and Lorenz. Other approaches such as the use of cryo-electron microscopy and synchrotron radiation have recently allowed increasing resolution and better understanding of the nature of the interactions and conformational changes implicated in the formation of actin filaments.[15][16]

Structure

Actin is one of the most abundant proteins in eukaryotes, where it is found throughout the cytoplasm.[3] In fact, in muscle fibres it comprises 20% of total cellular protein by weight and between 1% and 5% in other cells. However, there is not only one type of actin, the genes that code for actin are defined as a gene family (a family that in plants contains more than 60 elements, including genes and pseudogenes and in humans more than 30 elements).[17] This means that the genetic information of each individual contains instructions that generate actin variants (called isoforms) that possess slightly different functions. This, in turn, means that eukaryotic organisms express different genes that give rise to: α-actin, which is found in contractile structures; β-actin, found at the expanding edge of cells that use the projection of their cellular structures as their means of mobility; and γ-actin, which is found in the filaments of stress fibres.[18] In addition to the similarities that exist between an organism’s isoforms there is also an evolutionary conservation in the structure and function even between organisms contained in different eukaryotic domains: in bacteria the actin homologue MreB has been identified, which is a protein that is capable of polymerizing into microfilaments;[16] and in archaea the homologue Ta0583 is even more similar to the eukaryotic actins.[19]

Cellular actin has two forms: monomeric globules called G-actin and polymeric filaments called F-actin (that is, as filaments made up of many G-actin monomers). F-actin can also be described as a microfilament. Two parallel F-actin strands must rotate 166 degrees to lay correctly on top of each other. This creates the double helix structure of the microfilaments found in the cytoskeleton. Microfilaments measure approximately 7 nm in diameter with the helix repeating every 37 nm.Each strand of actin is bound to a molecule of adenosine triphosphate (ATP) or adenosine diphosphate (ADP) that is associated with a Mg2+ cation. The most commonly found forms of actin, compared to all the possible combinations, are ATP-G-Actin and ADP-F-actin.[20][21]

G-Actin

Scanning electron microscope images indicate that G-actin has a globular structure; however, X-ray crystallography shows that these globules are comprised of two lobes separated by a cleft. This structure represents the “ATPase fold”, which is a centre of enzymatic catalysis that binds ATP and Mg2+ and hydrolyzes the former to ADP plus phosphate. This fold is a conserved structural motif that is also found in other proteins that interact with triphosphate nucleotides such as hexokinase (an enzyme used in energy metabolism) or in Hsp70 proteins (a protein family that play an important part in protein folding).[22] G-actin is only functional when it contains either ADP or ATP in its cleft, however, the form that is bound to ATP predominates in cells when actin is present in its free state.[20]

Ribbon model of actin extracted from the striated muscle tissue of a rabbit after Graceffa and Domínguez, 2003. The four subdomains can be seen, as well as the N and C termini and the position of the ATP bond. The molecule is orientated using the usual convention of placing the - end (pointed end) in the upper part and the + end (barbed end) in the lower part.[12]

The X-ray crystallography model of actin that was produced by Kabsch from the striated muscle tissue of rabbits is the most commonly used in structural studies as it was the first to be purified. The G-actin crystallized by Kabsch is approximately 67 x 40 x 37 Å in size, has a molecular mass of 41,785 Da and an estimated isoelectric point of 4.8. Its net charge at pH = 7 is -7.[23] [24]

Primary structure

Elzinga and co-workers first determined the complete peptide sequence for this type of actin in 1973, with later work by the same author adding further detail to the model. It contains 374 amino acid residues. Its N-terminus is highly acidic and starts with an acetyled aspartate in its amino group. While its C-terminus is alkaline and is formed by a phenylalanine preceded by a cysteine, which has a degree of functional importance. Both extremes are in close proximity within the I-subdomain. An anomalous Nτ-methylhistidine is located at position 73.[23]

Tertiary structure - domains

The tertiary structure is formed by two domains known as the large and the small, which are separated by a cleft centred around the location of the bond with ATP-ADP+Pi. Below this there is a deeper notch called a “groove”. In the native state, despite their names, both have a comparable depth.[9]

The normal convention in topological studies means that a protein is shown with the biggest domain on the left-hand side and the smallest domain on the right-hand side. In this position the smaller domain is in turn divided into two: subdomain I (lower position, residues 1-32, 70-144 and 338-374) and subdomain II (upper position, residues 33-69). The larger domain is also divided in two: subdomain III (lower, residues 145-180 and 270-337) and subdomain IV (higher, residues 181-269). The exposed areas of subdomains I and III are referred to as the “barbed” ends, while the exposed areas of domains II and IV are termed the “pointed" ends. This nomenclature refers to the fact that, due to the small mass of subdomain II actin is polar; the importance of this will be discussed below in the discussion on assembly dynamics. Some authors call the subdomains Ia, Ib, IIa and IIb, respectively.[25]

Other important structures

The most notable supersecondary structure is a five chain beta sheet that is comprised of a β-meander and a β-α-β clockwise unit. It is present in both domains suggesting that the protein arose from gene duplication.[10]

  • The adenosine nucleotide binding site is located between two beta hairpin-shaped structures pertaining to the I and III domains. The residues that are involved are Asp11-Lys18 and Asp154-His161 respectively.
  • The divalent cation binding site is located just below that for the adenosine nucleotide. In vivo it is most often formed by Mg2+ or Ca2+ while in vitro it is formed by a chelating structure made up of Lys18 and two oxygens from the nucleotide’s α-and β-phosphates. This calcium is coordinated with six water molecules that are retained by the amino acids Asp11, Asp154, and Gln137. They form a complex with the nucleotide that restricts the movements of the so-called “hinge” region, located between residues 137 and 144. This maintains the native form of the protein until its withdrawal denatures the actin monomer. This region is also important because it determines whether the protein’s cleft is in the “open” or "closed” conformation.[25][12]
  • It is highly likely that there are at least three other centres with a lesser affinity (intermediate) and still others with a low affinity for divalent cations. It has been suggested that these centres may play a role in the polymerization of actin by acting during the activation stage.[25]
  • There is a structure in subdomain 2 that is called the “D-loop” because it binds with DNase I, it is located between the His40 and Gly48 residues. It has the appearance of a disorderly element in the majority of crystals, but it looks like a β-sheet when it is complexed with DNase I. Domínguez “et al.” suggest that the key event in polymerization is probably the propagation of a conformational change from the centre of the bond with the nucleotide to this domain, which changes from a loop to a spiral. However, this theory has been refuted by other studies.[12][26]

F-Actin

The classical description of F-actin states that is has a filamentous structure that can be considered to be a single stranded levorotatory helix with a rotation of 166º around the helical axis and an axial translation of 27.5 Å, or a single stranded dextrorotatory helix with a cross over spacing of 350-380 Å, with each actin surrounded by four more.[27] The symmetry of the actin polymer at 2.17 subunits per turn of a helix is incompatible with the formation of crystals, which is only possible with a symmetry of exactly 2, 3, 4 or 6 subunits per turn. Therefore, models have to be constructed that explain these anomalies using data from electron microscopy, cryo-electron microscopy, crystallization of dimers in different positions and diffraction of X-rays.[16] It should be pointed out that it is not correct to talk of a “structure” for a molecule as dynamic as the actin filament. In reality we talk of distinct structural states, in these the measurement of axial translation remains constant at 27.5 Å while the subunit rotation data shows considerable variability, with displacements of up to 10% from its optimum position commonly seen. Some proteins, such as cofilin appear to increase the angle of turn, but again this could be interpreted as the establishment of different "structural states". These could be important in the polymerization process.[28]

There is less agreement regarding measurements of the turn radius and filament thickness: while the first models assigned a longitude of 25 Å, current X-ray diffraction data, backed up by cryo-electron microscopy suggests a longitude of 23.7 Å. These studies have shown the precise contact points between monomers. Some are formed with units of the same chain, between the "barbed" end on one monomer and the "pointed" end of the next one. While the monomers in adjacent chains make lateral contact through projections from subdomain IV, with the most important projections being those formed by the C-terminus and the hydrophobic link formed by three bodies involving residues 39-42, 201-203 and 286. This model suggests that a filament is formed by monomers in a "sheet" formation, in which the subdomains turn about themselves, this form is also found in the bacterial actin homologue MreB.[16]

The F-actin polymer is considered to have structural polarity due to the fact that all the microfilament’s subunits point towards the same end. This gives rise to a naming convention: the end that possesses an actin subunit that has it’s ATP binding site exposed is called the «(-) end», while the opposite end where the cleft is directed at a different adjacent monomer is called the «(+) end».[18] The terms «pointed» and «barbed» referring to the two ends of the microfilaments derive from their appearance under transmission electron microscopy when samples are examined following a preparation technique called «decoration». This method consists of the addition of myosin S1 fragments to tissue that has been fixed with tannic acid. This myosin forms polar bonds with actin monomers, giving rise to a configuration that looks like arrows with feather fletchings along its shaft, where the shaft is the actin and the fletchings are the myosin. Following this logic, the end of the microfilament that does not have any protruding myosin is called the point of the arrow (- end) and the other end is called the barbed end (+ end).[29] A S1 fragment is composed of the head and neck domains of myosin II. Under physiological conditions, G-actin (the monomer form) is transformed to F-actin (the polymer form) by ATP, where the role of ATP is essential.[30]


The helical F-actin filament found in muscles also contains a tropomyosin molecule, which is a 40 nanometre long protein that is wrapped around the F-actin helix. During the resting phase the tropomyosin covers the actin’s active sites so that the actin-myosin interaction cannot take place and produce muscular contraction. There are other protein molecules bound to the tropomyosin thread, these are the troponins that have three polymers: troponin I, troponin T and troponin C.[31]

Folding

Ribbon model obtained using the PyMOL programme on cristalographs of the prefoldin proteins found in the arquea microorganism Pyrococcus horikoshii. The six supersecondary structures are present in a coiled helix “hanging” from the central beta barrels. These are often compared in the literature to the tentacles of a jellyfish. As far as is visible using electron microscopy, eukariotic prefoldin has a similar structure.[32]

Actin can spontaneously acquire a large part of its tertiary structure.[33] However, the way it acquires its fully functional form from its newly synthesized native form is special and almost unique in protein chemistry. The reason for this special route could be the need to avoid the presence of incorrectly folded actin monomers, which could be toxic as they can act as inefficient polymerization terminators. Nevertheless, it is key to establishing the stability of the cytoskeleton, and additionally, it is an essential process for coordinating the cell cycle.[34][35]

CCT is required in order to ensure that folding takes place correctly. CCT is a group II cytosolic molecular chaperone (or chaperonin, a protein that assists in the folding of other macromolecular structures). CCT is formed of a double ring of eight different subunits (hetero-octameric) and it differs from other molecular chaperones, particularly from its homologue GroEL which is found in arqueas, as it does not require a co-chaperone to act as a lid over the central catalytic cavity. Substrates bind to CCT through specific domains. It was initially thought that it only bound with actin and tubulin, although recent immunoprecipitation studies have shown that it interacts with a large number of polypeptides, which possibly function as substrates. It acts through ATP-dependent conformational changes that on occasion require several rounds of liberation and catalysis in order to complete a reaction.[36]

In order to successfully complete their folding, both actin and tubulin need to interact with another protein called prefoldin, which is a heterohexameric complex (formed by six distinct subunits), in an interaction that is so specific that the molecules have coevolved. Actin complexes with prefoldin while it is still being formed, when it is approximately 145 amino acids long, specifically those at the N-terminal.[37]

Different recognition sub-units are used for actin or tubulin although there is some overlap. In actin the subunits that bind with prefoldin are probably PFD3 and PFD4, which bind in two places one between residues 60-79 and the other between residues 170-198. The actin is recognized, loaded and delivered to the cytosolic chaperonin (CCT) in an open conformation by the inner end of prefoldin’s "tentacles” (see the image and note).[n. 1] The contact when actin is delivered is so brief that a tertiary complex is not formed, immediately freeing the prefoldin.[32]

Ribbon model of the apical γ-domain of the chaperonin CCT.

The CCT then causes actin's sequential folding by forming bonds with its subunits rather than simply enclosing it in its cavity.[n. 2] This is why it possesses specific recognition areas in its apical β-domain. The first stage in the folding consists of the recognition of residues 245-249. Next, other determinants establish contact.[38] Both actin and tubulin bind to CCT in open conformations in the absence of ATP. In actin’s case, two subunits are bound during each conformational change, whereas for tubulin binding takes place with four subunits. Actin has specific binding sequences, which interact with the δ and β-CCT subunits or with δ-CCT and ε-CCT. After AMP-PNP is bound to CCT the substrates move within the chaperonin’s cavity. It also seems that in the case of actin, the CAP protein is required as a possible cofactor in actin's final folding states.[35]

The exact manner by which this process is regulated is still not fully understood, but it is known that the protein PhLP3 (a protein similar to phosducin) inhibits its activity through the formation of a tertiary complex.[36]

ATPase’s catalytic mechanism

Actin is an ATPase, which means that it is an enzyme that hydrolyzes ATP. This group of enzymes is characterised by their slow reaction rates. It is know that this ATPase is “active”, that is, its speed increases by some 40,000 times when the actin forms part of a filament.[28] A reference value for this rate of hydrolysis under ideal conditions is around 0.3 s-1. Then, the Pi remains bound to the actin next to the ADP for a long time, until it is liberated next to the end of the filament.[39]

The exact molecular details of the catalytic mechanism are still not fully understood. Although there is much debate on this issue, it seems certain that a "closed" conformation is required for the hydrolysis of ATP, and it is thought that the residues that are involved in the process move to the appropriate distance.[28] The glutamic acid Glu137 is one of the key residues, which is located in subdomain 1. Its function is to bind the water molecule that produces a nucleophilic attack on the ATP’s γ-phosphate bond, while the nucleotide is strongly bound to subdomains 3 and 4. The slowness of the catalytic process is due to the large distance and skewed position of the water molecule in relation to the reactant. It is highly likely that the conformational change produced by the rotation of the domains between actin’s G and F forms moves the Glu137 closer allowing its hydrolysis. This model suggests that the polymerization and ATPase’s function would be decoupled straight away.[16]

Genetics

Principal interactions of structural proteins are at cadherin-based adherens junction. Actin filaments are linked to α-actinin and to the membrane through vinculin. The head domain of vinculin associates to E-cadherin via α-, β-, and γ-catenins. The tail domain of vinculin binds to membrane lipids and to actin filaments.

Actin has been one of the most highly conserved proteins throughout evolution because it interacts with a large number of other proteins. It has 80.2% sequence conservation at the gene level between Homo sapiens and Saccharomyces cerevisiae (a species of yeast), and 95% conservation of the primary structure of the protein product.

Although most yeasts have only a single actin gene, higher eukaryotes, in general, express several isoforms of actin encoded by a family of related genes. Mammals have at least six actin isoforms coded by separate genes,[40] which are divided into three classes (alpha, beta and gamma) according to their isoelectric points. In general, alpha actins are found in muscle (α-skeletal, α-aortic smooth, α-cardiac, and γ2-enteric smooth), whereas beta and gamma isoforms are prominent in non-muscle cells (β- and γ1-cytoplasmic). Although the amino acid sequences and in vitro properties of the isoforms are highly similar, these isoforms cannot completely substitute for one another in vivo.[41]

The typical actin gene has an approximately 100-nucleotide 5' UTR, a 1200-nucleotide translated region, and a 200-nucleotide 3' UTR. The majority of actin genes are interrupted by introns, with up to six introns in any of 19 well-characterised locations. The high conservation of the family makes actin the favoured model for studies comparing the introns-early and introns-late models of intron evolution.

All non-spherical prokaryotes appear to possess genes such as MreB, which encode homologues of actin; these genes are required for the cell's shape to be maintained. The plasmid-derived gene ParM encodes an actin-like protein whose polymerized form is dynamically unstable, and appears to partition the plasmid DNA into its daughter cells during cell division by a mechanism analogous to that employed by microtubules in eukaryotic mitosis.[42] Actin is found in both smooth and rough endoplasmic reticulums.


Assembly dynamics

Thin filament formation showing the polymerization mechanism for converting G-actin to F-actin; note the hydrolysis of the ATP.

Nucleation and polymerization

Actin polymerization and depolymerization is necessary in chemotaxis and cytokinesis. Nucleating factors are necessary to stimulate actin polymerization. One such nucleating factor is the Arp2/3 complex, which mimics a G-actin dimer in order to stimulate the nucleation of G-actin (or monomeric actin). The Arp2/3 complex binds to actin filaments at 70 degrees to form new actin branches off existing actin filaments. Also, actin filaments themselves bind ATP, and hydrolysis of this ATP stimulates destabilization of the polymer.

The growth of actin filaments can be regulated by thymosin and profilin. Thymosin binds to G-actin to buffer the polymerizing process, while profilin binds to G-actin to exchange ADP for ATP, promoting the monomeric addition to the barbed, plus end of F-actin filaments.

F-actin is both strong and dynamic. Unlike other polymers, such as DNA, whose constituent elements are bound together with covalent bonds, the monomers of actin filaments are assembled by weaker bonds. The lateral bonds with neighbouring monomers resolve this anomaly, which in theory should weaken the structure as they can be broken by thermal agitation. In addition, the weak bonds give the advantage that the filament ends can easily release or incorporate monomers. This means that the filaments can be rapidly remodelled and can change cellular structure in response to an environmental stimulus. Which, along with the biochemical mechanism by which it is brought about is known as the "assembly dynamic".[4]

In vitro studies

Studies focussing on the accumulation and loss of subunits by microfilaments are carried out in vitro (that is, in the laboratory and not on cellular systems) as the polymerization of the resulting actin gives rise to the same F-actin as produced in vivo. The in vivo process is controlled by a multitude of proteins in order to make it responsive to cellular demands, this makes it difficult to observe its basic conditions.[43] In vitro production takes place in a sequential manner: first, there is the «activation phase», when the bonding and exchange of divalent cations occurs in specific places on the G-actin, which is bound to ATP, this produces a conformational change, sometimes called G*-actin or F-actin monomer as it is very similar to the units that are located on the filament.[25] This prepares it for the «nucleation phase», in which the G-actin gives rise to small unstable fragments of F-actin that are able to polymerize. Unstable dimers and trimers are initially formed. The «elongation phase» begins when there are a sufficiently large number of these short polymers. In this phase the filament forms and rapidly grows through the reversible addition of new monomers at both extremes.[44] Finally, a «stationary equilibrium» is achieved where the G-actin monomers are exchanged at both ends of the microfilament without any change to its total length.[3] In this last phase the «critical concentration Cc» is defined as the ratio between the assembly constant and the dissociation constant for G-actin, where the dynamic for the addition and elimination of dimers and trimers does not produce a change in the microfilament's length. Under normal “in vitro” conditions Cc is 0.1 μM,[45] which means that at higher values polymerization occurs and at lower values depolymerization occurs.[46]

Role of ATP hydrolysis

As indicated above, although actin hydrolyzes ATP, everything points to the fact that ATP is not required for actin to be assembled, given that, on one hand, the hydrolysis mainly takes place inside the filament, and on the other hand the ADP could also instigate polymerization. This poses the question of understanding which thermodynamically unfavourable process requires such a prodigious expenditure of energy. The so-called “actin cycle”, which couples ATP hydrolysis to actin polymerization, consists of the preferential addition of G-actin-ATP monomers to a filament’s barbed end, and the simultaneous disassembly of F-actin-ADP monomers at the pointed end where the ADP is subsequently changed into ATP, thereby closing the cycle, this aspect of actin filament formation is known as “treadmilling”.

ATP is hydrolysed relatively rapidly just after the addition of a G-actin monomer to the filament. There are two hypotheses regarding how this occurs; the stochastic, which suggests that hydrolysis randomly occurs in a manner that is in some way influenced by the neighbouring molecules; and the vectoral, which suggests that hydrolysis only occurs adjacent to other molecules whose ATP has already been hydrolysed. In either case, the resulting Pi is not released, it remains for some time noncovalently bound to actin’s ADP, in this way there are three species of actin in a filament: ATP-Actin, ADP+Pi-Actin and ADP-Actin.[39] The amount of each one of these species present in a filament depends on its length and state: as elongation commences the filament has an approximately equal amount of actin monomers bound with ATP and ADP+Pi and a small amount of ADP-Actin at the (-) end. As the stationary state is reached the situation reverses, with ADP present along the majority of the filament and only the area nearest the (+) end containing ADP+Pi and with ATP only present at the tip.[47]

If we compare the filaments that only contain ADP-Actin with those that include ATP, in the former the critical constants are similar at both ends, while Cc for the other two nucleotides is different: At the (+) end Cc+=0.1 μM, while at the (-) end Cc-=0.8 μM, which gives rise to the following situations:[18]

  • For G-actin-ATP concentrations less than Cc+ no elongation of the filament occurs.
  • For G-actin-ATP concentrations less than Cc- but greater than Cc+ elongation occurs at the (+) end.
  • For G-actin-ATP concentrations greater than Cc- the microfilament grows at both ends.

It is therefore possible to deduce that the energy produced by hydrolysis is used to create a true “stationary state”, that is a flux, instead of a simple equilibrium, one that is dynamic, polar and attached to the filament. This justifies the expenditure of energy as it promotes essential biological functions.[39] In addition, the configuration of the different monomer types is detected by actin binding proteins, which also control this dynamism, as will be described in the following section.

Microfilament formation by treadmilling has been found to be atypical in stereocilia. In this case the control of the structure's size is totally apical and it is controlled in some way by gene expression, that is, by the total quantity of protein monomer synthesized in any given moment.[48]

Associated proteins

An actin (green) - profilin (blue) complex.[49] The profilin shown belongs to group II, normally present in the kidneys and the brain.

The actin cytoskeleton in vivo is not exclusively composed of actin, other proteins are required for its formation, continuance and function. These proteins are called actin-binding proteins' (ABP) and they are involved in actin’s polymerization, depolymerization, stability, organisation in bundles or networks, fragmentation and destruction.[3] The diversity of these proteins is such that actin is thought to be the protein that takes part in the greatest number of protein-protein interactions.[50] For example, G-actin sequestering elements exist that impede its incorporation into microfilaments. There are also proteins that stimulate its polymerization or that give complexity to the synthesizing networks.[18]

  • Thymosin β-4 is a 5 kDa protein that can bind with G-actin-ATP in a 1:1 stoichiometry; which means that one unit of thymosin β-4 binds to one unit of G-actin. Its role is to impede the incorporation of the monomers into the growing polymer.[51]
  • Profilin, is a cytosolic protein with a molecular weight of 15 kDa, which also binds with G-actin-ATP with a stoichiometry of 1:1, but it has a different function as it facilitates the replacement of ATP nucleotides by ADP. It is also implicated in other cellular functions, such as the binding of prolin repetitions in other proteins or of lipids that act as secondary messengers.[52][53]
The protein gelsolin, which is a key regulator in the assembly and disassembly of actin. It has six subdomains, S1-S6, each of which is composed of a five-stranded β-sheet flanked by two α-helices, one positioned perpendicular to the strands and the other in a parallel position. Both the N-terminal end, (S1-S3), and the C-terminal end, (S4-S6), form an extended β-sheet.[54]

Other proteins that bind to actin regulate the length of the microfilaments by cutting them, which gives rise to new active ends for polymerization. So that, if a microfilament, that has two ends that monomers can be added to or taken away from, is cut twice, there will be three new microfilaments with six ends. This new situation will favour the dynamics of assembly and disassembly. The most notable of these proteins are gelsolin and cofilin. These proteins first achieve a cut by binding to an actin monomer located in the polymer they then change the actin monomer’s conformation while remaining bound to the newly generated (+) end. This has the effect of impeding the addition or exchange of new G-actin subunits. Depolymerization is encouraged as the (-) ends are not linked to any other molecule.[55]

Other proteins that bind with actin cover the ends of F-actin in order to stabilize them, but they are unable to break them. Examples of this type of protein are CapZ (that binds the (+) ends depending on a cell’s levels of Ca2+/calmodulin. These levels depend on the cell’s internal and external signals and are involved in the regulation of its biological functions).[56] Another example is tropomodulin (that binds to the (-) end). Tropomodulin basically acts to stabilize the F-actin present in the myofibrils present in muscle [sarcomere]]s, which are structures characterized by their great stability.[57]

Atomic structure of Arp2/3.[58] Each colour corresponds to a subunit: Arp3, orange; Arp2, sea blue (subunits 1 and 2 are not shown); p40, green; p34, light blue; p20, dark blue; p21, magenta; p16, yellow.

The Arp2/3 complex is widely found in all eukaryotic organisms.[59] It is comprised of seven subunits, some of which possess a topology that is clearly related to their biological function: two of the subunits, «ARP2» and «ARP3», have a structure similar to that of actin monomers. This homology allows both units to act as nucleation agents in the polymerization of G-actin and F-actin. This complex is also required in more complicated processes such as in establishing dendritic structures and also in anastomosis (the reconnection of two branching structures that had previously been joined, such as in blood vessels).[60]

Chemical inhibitors

File:Phalloidin.png
Chemical structure of phalloidin.

There are a number of toxins that interfere with actin’s dynamics, either by preventing it from polymerizing (latrunculin and cytochalasin D) or by stabilizing it (phalloidin):

  • Latrunculin is a toxin produced by sponges, it binds to G-actin preventing it from binding with microfilaments.[61]
  • Cytocalasin D, is an alkaloid produced by fungi, that binds to the (+) end of F-actin preventing the addition of new monomers.[62] Cytocalasin D has been found to disrupt actin’s dynamics in protein p53, which has been found to affect the protein’s activity in animals[63] and gravitropic effects in plants.[64]
  • Phalloidin, is a toxin that has been isolated from the death cap mushroom Amanita phalloides, it binds to the interface between adjacent actin monomers in the F-actin polymer, preventing its depolymerization.[62]

Functions and location

Actin forms microfilaments that are typically one of the most dynamic of the three subclasses of the eukaryotic cytoskeleton. This gives actin major functions in cells:

  • Formation of microfilaments to give mechanical support to cells, and provide trafficking routes through the cytoplasm to aid signal transduction
  • Cell motility in cells which undergo amoeboid motion using pseudopods (see actoclampin molecular motors) and phagocytosis, for example of bacteria by macrophages
  • In metazoan muscle cells, to be the scaffold on which myosin proteins generate force to support muscle contraction
  • In non-muscle cells, to be a track for cargo transport myosins (nonconventional myosins) such as myosin V and VI. Nonconventional myosins use ATP hydrolysis to transport cargo, such as vesicles and organelles, in a directed fashion much faster than diffusion. Myosin V walks towards the barbed end of actin filaments, while myosin VI walks toward the pointed end. Most actin filaments are arranged with the barbed end toward the cellular membrane and the pointed end toward the cellular interior. This arrangement allows myosin V to be an effective motor for the export of cargos, and myosin VI to be an effective motor for import.

The actin protein is found in both the cytoplasm and the cell nucleus.[65] Its location is regulated by cell membrane signal transduction pathways that integrate the stimuli that a cell receives stimulating the restructuring of the actin networks in response. In Dictyostelium, [[phospholipase D] has been found to intervene in inositol phosphate pathways.[66] Actin filaments are particularly stable and abundant in muscle fibres. Within the sarcomere (the basic morphological and physiological unit of muscle fibres) actin is present in both the I and A bands; myosin is also present in the latter.[67]

Cytoskeleton

Fluorescence micrograph showing F-actin (in green) in rat fibroblasts.

Microfilaments are involved in the movement of all mobile cells, including non-muscular types, and drugs that disrupt F-actin organization (such as the cytochalasins) affect the activity of these cells. Actin comprises 2% of the total amount of proteins in hepatocytes, 10% in fibroblasts, 15% in amoebas and up to 50-80% in activated platelets.[68] There are a number of different types of actin with slightly different structures and functions. This means that α-actin is found exclusively in muscle fibres, while types β and γ are found in other cells. In addition, as the latter types have a high turnover rate the majority of them are found outside permanent structures. This means that the microfilaments found in cells other than muscle cells are present in two forms:[69]

  • Microfilament networks. Animal cells commonly have a cell cortex under the cell membrane that contains a large number of microfilaments, which precludes the presence of organelles. This network is connected with numerous receptor cells that relay signals to the outside of a cell.
  • Microfilament bundles. These extremely long microfilaments are located in networks and, in association with contractile proteins such as non-muscular myosin, they are involved in the movement of substances at an intracellular level.

Yeasts

Actin’s cytoskeleton is key to the processes of endocytosis, cytokinesis, determination of cell polarity and morphogenesis in yeasts. In addition to relying on actin these processes involve 20 or 30 associated proteins, which all have a high degree of evolutionary conservation, along with many signalling molecules. Together these elements allow a spatially and temporally modulated assembly that defines a cell’s response to both internal and external stimuli.[70]

Yeasts contain three main elements that are associated with actin: patches, cables and rings that, despite being present for long periods of time, are subject to a dynamic equilibrium due to continual polymerization and depolymerization. They possess a number of accessory proteins including ADF/cofilin, which has a molecular weight of 16kDa and is coded for by a single gene, called COF1; Aip1, a cofilin cofactor that promotes the disassembly of microfilaments; Srv2/CAP, a process regulator related to adenylate cyclase proteins; a profilin with a molecular weight of approximately 14 kDa that is associated with actin monomers; and twinfilin, a 40 kDa protein involved in the organization of patches.[70]

Plants

Plant genome studies have revealed the existence of protein isovariants within the actin family of genes. Within Arabidopsis thaliana, a dicotyledon used as a model organism, there are ten types of actin, nine types of α-tubulins, six β-tubulins, six profilins and dozens of myosins. This diversity is explained by the evolutionary necessity of possessing variants that slightly differ in their temporal and spatial expression. The majority of these proteins were jointly expressed in the tissue analysed. Actin networks are distributed throughout the cytoplasm of cells that have been cultivated in vitro. There is a concentration of the network around the nucleus that is connected via spokes to the cellular cortex, this network is highly dynamic, with a continuous polymerization and depolymerization.[71]

Structure of the C-terminal subdomain of villin, a protein capable of splitting microfilaments.[72]

Even though the majority of plant cells have a cell wall that defines their morphology and impedes their movement, their microfilaments can generate sufficient force to achieve a number of cellular activities, such as, the cytoplasmic currents generated by the microfilaments and myosin. Actin is also involved in the movement of organelles and in cellular morphogenesis, which involve cell division as well as the elongation and differentiation of the cell.[73]

The most notable proteins associated with the actin cytoskeleton in plants include:[73] villin, which belongs to the same family as gelsolin/severin and is able to cut microfilaments and bind actin monomers in the presence of calcium cations; fimbrin, which is able to recognize and unite actin monomers and which is involved in the formation of networks (by a different regulation process from that of animals and yeasts);[74] formins, which are able to act as an F-actin polymerization nucleating agent; myosin, a typical molecular motor that is specific to eukaryotes and which in Arabidopsis thaliana is coded for by 17 genes in two distinct classes; CHUP1, which can bind actin and is implicated in the spatial distribution of chloroplasts in the cell; KAM1/MUR3 that define the morphology of the Golgi apparatus as well as the composition of xyloglucans in the cell wall; NtWLIM1, which facilitates the emergence of actin cell structures; and ERD10, which is involved in the association of organelles within membranes and microfilaments and which seems to play a role that is involved in an organism’s reaction to stress.

Nuclear actin

Actin is essential for transcription from RNA polymerases Pol I, Pol II and Pol III. In Pol I transcription, actin and myosin (MYO1C, which binds DNA) act as a molecular motor. For Pol II transcription, β-actin is needed for the formation of the preinitiation complex. Pol III contains β-actin as a subunit. Actin can also be a component of chromatin remodelling complexes as well as pre-mRNP particles (that is, precursor messenger RNA bundled in proteins), and is involved in nuclear export of RNAs and proteins.[75]


Muscular contraction

The structure of a sarcomere, the basic morphological and functional unit of the skeletal muscles that contains actin.

Outline of a muscle contraction

In muscle, actin is the major component of thin filaments, which, together with the motor protein myosin (which forms thick filaments), are arranged into actomyosin myofibrils. These fibrils comprise the mechanism of muscle contraction. Using the hydrolysis of ATP for energy, myosin heads undergo a cycle during which they attach to thin filaments, exert a tension, and then, depending on the load, perform a power stroke that causes the thin filaments to slide past, shortening the muscle.

In contractile bundles, the actin-bundling protein alpha-actinin separates each thin filament by ~35 nm. This increase in distance allows thick filaments to fit in between and interact, enabling deformation or contraction. In deformation, one end of myosin is bound to the plasma membrane, while the other end "walks" toward the plus end of the actin filament. This pulls the membrane into a different shape relative to the cell cortex. For contraction, the myosin molecule is usually bound to two separate filaments and both ends simultaneously "walk" toward their filament's plus end, sliding the actin filaments closer to each other. This results in the shortening, or contraction, of the actin bundle (but not the filament). This mechanism is responsible for muscle contraction and cytokinesis, the division of one cell into two.

Actin’s role in muscle contraction

The helical F-actin filament found in muscles also contains a tropomyosin molecule, which is a 40 nanometres long protein that is wrapped around the F-actin helix. During the resting phase the tropomyosin covers the actin’s active sites so that the actin-myosin interaction cannot take place and produce muscular contraction (the interaction gives rise to a movement between the two proteins that, because it is repeated many times, produces a contraction). There are other protein molecules bound to the tropomyosin thread, these include the troponins that have three polymers: troponin I, troponin T and troponin C.[31] Tropomyosin’s regulatory function depends on its interaction with troponin in the presence of Ca2+ ions.[76]

Both actin and myosin are involved in muscle contraction and relaxation and they make up 90% of muscle protein.[77] The overall process is initiated by an external signal, typically through an action potential stimulating the muscle, which contains specialized cells whose interiors are rich in actin and myosin filaments. The contraction-relaxation cycle comprises the following steps:[78]

  1. Depolarization of the sarcolemma and transmission of an action potential through the T-tubules.
  2. Opening of the sarcoplasmic reticulum’s Ca2+ channels.
  3. Increase in cytosolic Ca2+ concentrations and the interaction of these cations with troponin causing a conformational change in its structure. This in turn alters the structure of tropomyosin, which covers actin’s active site, allowing the formation of myosin-actin cross-links (the latter being present as thin filaments).[31]
  4. Movement of myosin heads over the thin filaments, this can either involve ATP or be independent of ATP. The former mechanism, mediated by ATPase activity in the myosin heads, causes the movement of the actin filaments towards the Z-disc.
  5. Ca2+ capture by the sarcoplasmic reticulum, causing a new conformational change in tropomyosin that inhibits the actin-myosin interaction.[77]

Other biological processes

The traditional image of actin’s function relates it to the maintenance of the cytoskeleton and, therefore, the organization and movement of organelles, as well as the determination of a cell’s shape.[69] However, actin has a wider role in eukaryotic cell physiology, in addition to similar functions in prokaryotes.

  • Cytokinesis. Cell division in animal cells and yeasts normally involves the separation of the parent cell into two daughter cells through the constriction of the central circumference. This process involves a constricting ring composed of actin, myosin and α-actinin.[79] In the "fission yeast” Schizosaccharomyces pombe, actin is actively formed in the constricting ring with the participation of Arp3, the formin Cdc12, profilin and WASp, along with preformed microfilaments. Once the ring has been constructed the structure is maintained by a continual assembly and disassembly that, aided by the Arp2/3 complex and formins, is key to one of the central processes of cytokinesis.[80] The totality of the contractile ring, the spindle apparatus, microtubules and the dense peripheral material is called the «Fleming body» or «intermediate body».[69]
  • Apoptosis. During programmed cell death the ICE/ced-3 family of proteases (one of the interleukin-1β-converter proteases) degrade actin into two fragments “in vivo”, one of the fragments is 15 kDa and the other 31 kDa. This represents one of the mechanisms involved in destroying cell viability that form the basis of apoptosis.[81] The protease calpain has also been shown to be involved in this type of cell destruction;[82] just as the use of calpain inhibitors has been shown to decrease actin proteolysis and the degradation of DNA (another of the characteristic elements of apoptosis).[83] On the other hand, the stress-induced triggering of apoptosis causes the reorganization of the actin cytoskeleton (which also involves its polymerization), giving rise to structures called stress fibres; this is activated by the MAP kinase pathway.[84]
Diagram of a zonula occludens or tight junction, a structure that joins the epithelium of two cells. Actin is one of the anchoring elements shown in green.
  • Cellular adhesion and development. The adhesion between cells is a characteristic of multicellular organisms that enables tissue specialization and therefore increases cell complexity. Adhesion of cell epithelia involves the actin cytoskeleton in each of the joined cells as well as cadherins acting as extracellular elements with the connection between the two mediated by catenins.[85] Interfering in actin dynamics has repercussions for an organism’s development, in fact actin is such a crucial element that systems of redundant genes are available. For example, if the α-actinin or gelation factor gene has been removed in Dictyostelium individuals do not show an anomalous phenotype possibly due to the fact that each of the proteins can perform the function of the other. However, the development of double mutations that lack both gene types is affected.[86]
  • Gene expression modulation. Actin’s state of polymerization affects the pattern of gene expression. In 1997, it was discovered that cytocalasin D-mediated depolymerization in Schwann cells causes a specific pattern of expression for the genes involved in the myelinization of this type of nerve cell.[87] F-actin has been shown to modify the transcriptome in some of the life stages of unicellular organisms, such as the fungus Candida albicans.[88] In addition, proteins that are similar to actin play a regulatory role during spermatogenesis in mice[89] and, in yeasts, actin-like proteins are thought to play a role in the regulation of gene expression.[90] In fact, actin is capable of acting as a transcription initiator when it reacts with a type of nuclear myosin that interacts with RNA polymerases and other enzymes involved in the transcription process.[65]
  • Stereocilia dynamics. Some cells develop fine filliform outgrowths on their surface that have a mechanosensory function. For example, this type of organelle is present in the Organ of Corti, which is located in the ear . The main characteristic of these structures is that their length can be modified.[91] The molecular architecture of the stereocilia includes a paracrystalline actin core in dynamic equilibrium with the monomers present in the adjacent cytosol. Type VI and VIIa myosins are present throughout this core, while myosin XVa is present in its extremities in quantities that are proportional to the length of the stereocilia.[92]

Molecular pathology

The majority of mammals posses six different actin genes. Of these, two code for the cytoskeleton (ACTB and ACTG1) while the other four are involved in skeletal striated muscle (ACTA1), smooth muscle tissue (ACTA2), intestinal muscles (ACTG2) and cardiac muscle (ACTC1). In this way the actin in the cytoskeleton is involved in the pathogenic mechanisms of many infectious agents, including HIV. The vast majority of the mutations that affect actin are point mutations that have a dominant effect, with the exception of six mutations involved in nemaline myopathy. This is because in many cases the mutant of the actin monomer acts as a “cap” by preventing the elongation of F-actin.[25]

Pathology associated with ACTA1

ACTA1 is the gene that codes for the α-isoform of actin that is predominant in human skeletal striated muscles, although it is also expressed in heart muscle and in the thyroid gland.[93] Its DNA sequence consists of seven exons that produce five known transcripts.[94] The majority of these consist of point mutations causing substitution of amino acids. The mutations are in many cases associated with a phenotype that determines the severity and the course of the affliction.[25][94]

Giant nemaline rods produced by the transfection of a DNA sequence of ACTA1, which is the carrier of a mutation responsible for nemaline myopathy.[95]

The mutation alters the structure and function of skeletal muscles producing one of three forms of myopathy: type 3 nemaline myopathy, congenital myopathy with an excess of thin myofilaments (CM) and Congenital myopathy with fibre type disproportion (CMFTD). Mutations have also been found that produce “core” myopathies).[96] Although their phenotypes are similar, in addition to typical nemaline myopathy some specialists distinguish another type of myopathy called actinic nemaline myopathy. In the former, clumps of actin form instead of the typical rods. It is important to state that a patient can show more than one of these phenotypes in a biopsy.[97] The most common symptoms consist of a typical facial morphology (myopathic faces), muscular weakness, a delay in motor development and respiratory difficulties. The course of the illness, its gravity and the age at which it appears are all variable and overlapping forms of myopathy are also found. A symptom of nemalinic myopathy is that “Nemaline rods” appear in differing places in Type 1 muscle fibres. These rods are non-pathognomonic structures that have a similar composition to the Z disks found in the sarcomere.[98]

The pathogenesis of this myopathy is very varied. Many mutations occur in the region of actin’s indentation near to its nucleotide binding sites, while others occur in Domain 2, or in the areas where interaction occurs with associated proteins. This goes some way to explain the great variety of clumps that form in these cases, such as Nemaline or Intranuclear Bodies or Zebra Bodies. [25] Changes in actin’s folding occur in nemaline myopathy as well as changes in its aggregation and there are also changes in the expression of other associated proteins. In some variants where intranuclear bodies are found the changes in the folding masks the nucleus’s protein exportation signal so that the accumulation of actin's mutated form occurs in the cell nucleus.[99] On the other hand it appears that mutations to ACTA1 that give rise to a CFTDM have a greater effect on sarcomeric function than on its structure.[100] Recent investigations have tried to understand this apparent paradox, which suggests there is no clear correlation between the number of rods and muscular weakness. It appears that some mutations are able to induce a greater apoptosis rate in type II muscular fibres.[34]

Position of seven mutations relevant to the various actinopathies related to ACTA1.[95]

In smooth muscle

There are two isoforms that code for actins in the smooth muscle tissue:

ACTG2 codes for the largest actin isoform, which has nine exons, one of which, the one located at the 5' end, is not translated.[101] It is an γ-actin that is expressed in the enteric smooth muscle. No mutations to this gene have been found that correspond to pathologies, although microarrays have shown that this protein is more often expressed in cases that are resistant to chemotherapy using cisplatin.[102]

ACTA2 codes for an α-actin located in the smooth muscle, and also in vascular smooth muscle. It has been noted that the MYH11 mutation could be responsible for at least 14% of hereditary thoracic aortic aneurisms particularly Type 6. This is because the mutated variant produces an incorrect filamentary assembly and a reduced capacity for vascular smooth muscle contraction. Degradation of the aortic media has been recorded in these individuals, with areas of disorganization and hyperplasia as well as stenosis of the aorta’s vasa vasorum.[103] The number of afflictions that the gene is implicated in is increasing. It has been related to Moyamoya disease and it seems likely that certain mutations in heterozygosis could confer a predisposition to many vascular pathologies, such as thoracic aortic aneurysm and ischaemic heart disease.[104] The α-actin found in smooth muscles is also an interesting marker for evaluating the progress of liver cirrhosis.[105]

In heart muscle

The ACTC1 gene codes for the α-actin isoform present in heart muscle. It was first sequenced by Hamada and co-workers in 1982, when it was found that it is interrupted by five introns.[106] It was the first of the six genes where alleles were found that were implicated in pathological processes.[107]

Crossection of a rat heart that is showing signs of dilated cardiomyopathy.[108]

A number of structural disorders associated with point mutations of this gene have been described that cause malfunctioning of the heart, such as Type 1R dilated cardiomyopathy and Type 11 hypertrophic cardiomyopathy. Certain defects of the atrial septum have been described recently that could also be related to these mutations.[109][110]

Two cases of dilated cardiomyopathy have been studied involving a substitution of highly conserved amino acids belonging to the protein domains that bind and intersperse with the Z discs. This has led to the theory that the dilation is produced by a defect in the transmission of contractile force in the myocytes.[107][27]

The mutations in”ACTC1” are responsible for at least 5% of hypertrophic cardiomyopathies.[111] The existence of a number of point mutations have also been found:[112]

  • Mutation E101K: changes of net charge and formation of a weak electrostatic link in the actomyosin-binding site.
  • P166A: interaction zone between actin monomers.
  • A333P: actin-myosin interaction zone.

Pathogenesis appears to involve a compensatory mechanism: the mutated proteins act like “toxins” with a dominant effect, decreasing the heart’s ability to contract causing abnormal mechanical behaviour such that the hypertrophy, that is usually delayed, is a consequence of the cardiac muscle’s normal response to stress.[113]

Recent studies have discovered “ACTC1” mutations that are implicated in two other pathological processes: Infantile idiopathic restrictive cardiomyopathy,[114] and noncompaction of the left ventricular myocardium.[115]

In cytoplasmatic actins

ACTB is a highly complex locus. A number of pseudogenes exist that are distributed throughout the genome, and its sequence contains six exons that can give rise to up to 21 different transcriptions by alternative splicing, which are known as the β-actins. Consistent with this complexity, its products are also found in a number of locations and they form part of a wide variety of processes (cytoskeleton, NuA4 histone-acyltransferase complex, cell nucleus) and in addition they are associated with the mechanisms of a great number of pathological processes (carcinomas, juvenile dystonia, infection mechanisms, nervous system malformations and tumour invasion, among others).[116] A new form of actin has been discovered, kappa actin, which appears to substitute for β-actin in processes relating to tumours.[117]

Image taken using confocal microscopy and employing the use of specific antibodies showing actin’s cortical network. In the same way that in juvenile dystonia there is an interruption in the structures of the cytoskeleton, in this case it is produced by cytochalasin D.[118]

Three pathological processes have so far been discovered that are caused by a direct alteration in gene sequence:

  • A dominant point mutation has also been discovered that causes neutrophil granulocyte dysfunction and recurring infections. It appears that the mutation modifies the domain responsible for binding between profilin and other regulatory proteins. Actin’s affinity for profilin is greatly reduced in this allele.[121]

The ACTG1 locus codes for the cytosolic γ-actin protein that is responsible for the formation of cytoskeletal microfilaments. It contains 6 exons, giving rise to 22 different mRNAs, which produce 4 complete isoforms whose form of expression is probably dependent on the type of tissue they are found in. It also has two different DNA promoters. [122] It has been noted that the sequences translated from this locus and from that of β-actin are very similar to the predicted ones, suggesting a common ancestral sequence that suffered duplication and genetic conversion.[123]

In terms of pathology, it has been associated with processes such as amyloidosis, retinitis pigmentosa, infection mechanisms, kidney diseases and various types of congenital hearing loss.[122]

Six autosomal-dominant point mutations in the sequence have been found to cause various types of hearing loss, particularly sensorineural hearing loss linked to the DFNA 20/26 locus. It seems that they affect the stereocilia of the ciliated cells present in the inner ear’s Organ of Corti. β-actin is the most abundant protein found in human tissue, but it is not very abundant in ciliated cells, which explains the location of the pathology. On the other hand, it appears that the majority of these mutations affect the areas involved in linking with other proteins, particularly actomyosin.[25] Some experiments have suggested that the pathological mechanism for this type of hearing loss relates to that fact that the F-actin in the mutations is more sensitive to cofilin than normal.[124]

However, although there is no record of any case, it is know that γ-actin is also expressed in skeletal muscles, and although it is present in small quantities, model organisms have shown that its absence can give rise to myopathys.[125]

Other pathological mechanisms

Some infectious agents use actin, especially cytoplasmic actin, in their life cycle. Two basic forms are present in bacteria:

  • Listeria monocytogenes, some species of Rickettsia, Shigella flexneri and other intracellular germs escape from phagocytic vacuoles by coating themselves with a capsule of actin filaments. L. monocytogenes and S. flexneri both generate a tail in the form of a "comet tail" that gives them mobility. Each species exhibits small differences in the molecular polymerization mechanism of their «comet tails». Different displacement velocities have been observed, for example, with Listeria and Shigella found to be the fastest.[126] Many experiments have demonstrated this mechanism in vitro. This indicates that the bacteria are not using a myosin-like protein motor, and it appears that their propulsion is acquired from the pressure exerted by the polymerization that takes place near to the microorganism's cell wall. The bacteria have previously been surrounded by ABPs from the host, and as a minimum the covering contains Arp2/3 complex, Ena/VASP proteins, cofilin, a buffering protein and nucleation promoters, such as vinculin complex. Through these movements they form protrusions that reach the neighbouring cells, infecting them as well so that the immune system can only fight the infection through cell immunity. The movement could be caused by the modification of the curve and debranching of the filaments.[127] Other species, such as Mycobacterium marinum and Burkholderia pseudomallei, are also capable of localized polymerization of cellular actin to aid their movement through a mechanism that is centered on the Arp2/3 complex. In addition the vaccine virus Vaccinia also uses elements of the actin cytoskeleton for its dissemination.[128]
  • Pseudomonas aeruginosa is able to form a protective biofilm in order to escape a host organism’s defences, especially white blood cells and antibiotics. The biofilm is constructed using DNA and actin filaments from the host organism.[129]


In addition to the previously cited example, actin polymerization is stimulated in the initial steps of the internalization of some viruses, notably HIV, by, for example, inactivating the cofilin complex.[130]

The role that actin plays in the invasion process of cancer cells has still not been determined.[131]

Evolution

The eukaryotic cytoskeleton of organisms among all taxonomic groups have similar components to actin and tubulin. For example, the protein that is coded by the ACTG2 gene in humans is completely equivalent to the homologues present in rats and mice, even though at a nucleotide level the similarity decreases to 92 %. [132] However, there are major differences with the equivalents in prokaryotes (FtsZ and MreB), where the similarity between nucleotide sequences is between 40−50 % among different bacteria and archaea species. Some authors suggest that the ancestral protein that gave rise to the model eukaryotic actin resembles the proteins present in modern bacterial cytoskeletons.[133]

Structure of MreB, a bacterial protein whose three-dimensional structure resembles that of G-actin.

Some authors point out that the behaviour of actin, tubulin and histone, a protein involved in the stabilization and regulation of DNA, are similar in their ability to bind nucleotides and in their ability of take advantage of Brownian motion. It has also been suggested that they all have a common ancestor.[134] Therefore evolutionary processes resulted in the diversification of ancestral proteins into the varieties present today, conserving, among others, actins as efficient molecules that were able to tackle essential ancestral biological processes, such as endocytosis.[135]

Equivalents in bacteria

The bacterial cytoskeleton may not be as complex as that found in eukaryotes, however, it contains proteins that are highly similar to actin monomers and polymers. The bacterial protein MreB polymerizes into thin non-helical filaments and occasionally into helical structures similar to F-actin. [16] Furthermore its crystalline structure is very similar to that of G-actin (in terms of its three dimensional conformation), there are even similarities between the MreB protofilaments and F-actin. The bacterial cytoskeleton also contains the FtsZ proteins, which are similar to tubulin.[136]

Bacteria therefore possess a cytoskeleton with homologous elements to actin (for example, MreB, ParM, and MamK), even though the amino acid sequence of these proteins diverges from that present in animal cells. However, MreB and ParM have a high degree of structural similarity to eukaryotic actin. The highly dynamic microfilaments formed by the aggregation of MreB and ParM are essential to cell viability and they are involved in cell morphogenesis, chromosome segregation and cell polarity. ParM is an actin homologue that is coded in a plasmid and it is involved in the regulation of plasmid DNA.[137]

Applications

Actin is used in scientific and technological laboratories as a track for molecular motors such as myosin (either in muscle tissue or outside it) and as a necessary component for cellular functioning. It can also be used as a diagnostic tool, as several of its anomalous variants are related to the appearance of specific pathologies.

  • Nanotechnology. Actin-myosin systems act as molecular motors that permit the transport of vesicles and organelles throughout the cytoplasm. It is possible that actin could be applied to nanotechnology as its dynamic ability has been harnessed in a number of experiments including those carried out in acellular systems. The underlying idea is to use the microfilaments as tracks to guide molecular motors that can transport a given load. That is actin could be used to define a circuit along which a load can be transported in a more or less controlled and directed manner. In terms of general applications, it could be used for the directed transport of molecules for deposit in determined locations, which would permit the controlled assembly of nanostructures.[138] These attributes could be applied to laboratory processes such as on lab-on-a-chip, in nanocomponent mechanics and in nanotransformers that convert mechanical energy into electrical energy.[139]
  • Internal control of techniques used in molecular biology, such as western blot and real-time polymerase chain reaction. As actin is essential for cell survival it has been postulated that the quantity of actin is under such tight control at a cellular level that it can be assumed that its transcription (that is, the degree to which its genes are expressed) and translation, that is the production of protein, is practically constant and independent of experimental conditions. Therefore, it is common practice in protein quantification studies (western blot) and transcription studies (Real-time polymerase chain reaction) to carry out the quantification of the gene of interest and also the quantification of a reference gene such as the one that codes for actin. By dividing the quantity of the gene of interest by that of the actin gene it is possible to obtain a relative quantity that can be compared between different experiments,[140] whenever the expression of the latter is constant. It is worth pointing out that actin does not always have the desired stability in its gene expression.[141]
  • Health. Some alleles of actin cause diseases, for this reason techniques for their detection have been developed. In addition, actin can be used as an indirect marker in surgical pathology: it is possible to use variations in the pattern of its distribution in tissue as a marker of invasion in neoplasia, vasculitis and other conditions.[142] Further, due to actin’s close association with the apparatus of muscular contraction its levels in skeletal muscle diminishes when these tissues atrophy, it can therefore be used as a marker of this physiological process. [143]
  • Food technology. It is possible to determine the quality of certain processed foods, such as sausages, by quantifying the amount of actin present in the constituent meat. Traditionally, a method has been used that is based on the detection of 3-methylhistidine in hydrolyzed samples of these products, as this compound is present in actin and F-myosin’s heavy chain (both are major components of muscle). The generation of this compound in animal flesh derives from the methylation of histidine residues present in both proteins.[144][145]

See also

Notes

  1. ^ Holmes; Popp; Gebhard; Kabsch (1990). "Atomic model of the actin filament" (w). 347 (6288): 44–49. doi:10.1038/347044a0. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |name4= ignored (help); Unknown parameter |pub-journal= ignored (help)
  2. ^ Doherty GJ and McMahon HT (2008). "Mediation, Modulation and Consequences of Membrane-Cytoskeleton Interactions". Annual Review of Biophysics. 37: 65–95. doi:10.1146/annurev.biophys.37.032807.125912. PMID 18573073.
  3. ^ a b c d Marc Maillet Biología celular (in Spanish). Published by Elsevier España, 2002; page 132. ISBN 84-458-1105-3
  4. ^ a b Alberts; et al. (2004). Biología molecular de la célula (in Spanish). Barcelona: Omega. ISBN 54-282-1351-8. {{cite book}}: Explicit use of et al. in: |author= (help)
  5. ^ Halliburton, W.D. (1887). "On muscle plasma". J. Physiol. 8 (3–4): 133. PMC 1485127. PMID 16991477.
  6. ^ Szent-Gyorgyi, A. (1945). "Studies on muscle". Acta Physiol Scandinav. 9 (Suppl): 25.
  7. ^ a b Straub FB, Feuer G (1989). "Adenosine triphosphate. The functional group of actin. 1950". Biochim. Biophys. Acta. 1000: 180–95. PMID 2673365. Cite error: The named reference "Straub" was defined multiple times with different content (see the help page).
  8. ^ Bárány M, Barron JT, Gu L, Bárány K (2001). "Exchange of the actin-bound nucleotide in intact arterial smooth muscle". J. Biol. Chem. 276 (51): 48398–403. doi:10.1074/jbc.M106227200. PMID 11602582. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  9. ^ a b Elzinga; Collins; Kuehl; Adelstein (1973). "Complete amino-acid sequence of actin of rabbit skeletal muscle" (PDF). 70 (9): 2687–2691. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |access date= ignored (|access-date= suggested) (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help)
  10. ^ a b Kabsch W, Mannherz HG, Suck D, Pai EF, Holmes KC (1990). "Atomic structure of the actin:DNase I complex". Nature. 347 (6288): 37–44. doi:10.1038/347037a0. PMID 2395459. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  11. ^ Holmes KC, Popp D, Gebhard W, Kabsch W (1990). "Atomic model of the actin filament". Nature. 347 (6288): 44–9. doi:10.1038/347044a0. PMID 2395461. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  12. ^ a b c d Domínguez (2001). "The crystal structure of uncomplexed actin in the ADP state". 293 (5530): 708–11. PMID 11474115. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help); Unknown parameter |name= ignored (help)
  13. ^ Oriol (1977). "Crystallization of native striated-muscle actin". FEBS Lett. 73 (1): 89–91. PMID 320040. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help); Unknown parameter |name= ignored (help)
  14. ^ Sawaya MR, Kudryashov DS, Pashkov I, Adisetiyo H, Reisler E, Yeates TO. (2008). "Multiple crystal structures of actin dimers and their implications for interactions in the actin filament". Acta Crystallogr D Biol Crystallogr. PMID 18391412.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  15. ^ Narita (2006). "Structural basis of actin filament capping at the barbed-end: a cryo-electron microscopy study" (PDF). Embo J. 25 (23): 5626–33. {{cite journal}}: Text "doi 10.1038/sj.emboj.7601395" ignored (help); Unknown parameter |access date= ignored (|access-date= suggested) (help)
  16. ^ a b c d e f Toshiro Oda, Mitsusada Iwasa, Tomoki Aihara, Yuichiro Maéda and Akihiro Narita (2009): The nature of the globular- to fibrous-actin transition. Nature 457, 441-445 (22 January 2009) | doi:10.1038/nature07685
  17. ^ Ponte; Gunning; Blau; Kedes (1983). "Human actin genes are single copy for alpha-skeletal and alpha-cardiac actin but multicopy for β- and γ-cytoskeletal genes: 3' untranslated regions are isotype specific but are conserved in evolution". 3 (10): 1783–1791. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |name4= ignored (help); Unknown parameter |pub-journal= ignored (help)
  18. ^ a b c d Lodish; et al. (2005). Biología celular y molecular (in Spanish). Buenos Aires: Médica Panamericana. ISBN 950-06-1974-3. {{cite book}}: Explicit use of et al. in: |author= (help)
  19. ^ Futoshi Hara, Kan Yamashiro, Naoki Nemoto, Yoshinori Ohta, Shin-ichi Yokobori, Takuo Yasunaga, Shin-ichi Hisanaga, and Akihiko Yamagishi. (2007): An Actin Homologue of the Archaeon Thermoplasma acidophilum That Retains the Ancient Characteristics of Eukaryotic Actin. Journal of Bacteriology, p. 2039-2045, Vol. 189, No. 5 doi:10.1128/JB.01454-06
  20. ^ a b Graceffa; Dominguez (2003). "Crystal Structure of Monomeric Actin in the ATP State: STRUCTURAL BASIS OF NUCLEOTIDE-DEPENDENT ACTIN DYNAMICS". 278 (36): 34172–34180. doi:10.1074/jbc.M303689200. PMID 12813032. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |pub-journal= ignored (help)CS1 maint: unflagged free DOI (link)
  21. ^ Reisler (1993). "Actin molecular structure and function". 5 (1): 41–7. doi:10.1016/S0955-0674(05)80006-7. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name= ignored (help); Unknown parameter |pub-journal= ignored (help)
  22. ^ "NCBI Conserved Domains: ATP binding site". {{cite web}}: Unknown parameter |access date= ignored (|access-date= suggested) (help); Unknown parameter |access year= ignored (help)
  23. ^ a b Elzinga (1975). "The primary structure of actin from rabbit skeletal muscle. Completion and analysis of the amino acid sequence". J Biol Chem. 250 (15). {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help)
  24. ^ Elzinga. "Complete amino-acid sequence of actin of rabbit skeletal muscle" (PDF). Proc Natl Acad Sci U S A. 70 (9). {{cite journal}}: Unknown parameter |co-authors= ignored (help)
  25. ^ a b c d e f g h Cristóbal G. Dos Remedios, Deepak Chhabra (2008). Actin-binding Proteins and Disease. Springer. ISBN 0-387-71747-1. See in Google books
  26. ^ Rould (2006). "Crystal structures of expressed non-polymerizable monomeric actin in the ADP and ATP states". J Biol chem. 281 (42). doi:10.1074/jbc.M601973200. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help)CS1 maint: unflagged free DOI (link)
  27. ^ a b "23". Bioquímica: Libro de texto con aplicaciones clínicas (in Spanish) (4 ed.). Reverte. 2004. p. 1021. ISBN 8429172084. {{cite book}}: Unknown parameter |name= ignored (help); Unknown parameter |surnames= ignored (help).
  28. ^ a b c Egelman (2007). "Actin structure and function: what we still do not understand". J Biol Chem. 282 (50). doi:10.1074/jbc.R700030200. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help)CS1 maint: unflagged free DOI (link)
  29. ^ DA Begg, R Rodewald and LI Rebhun (1978): The visualization of actin filament polarity in thin sections. Evidence for the uniform polarity of membrane-associated filaments. The Journal of Cell Biology, Vol 79, 846-852.
  30. ^ Histologi by Finn Geneser p. 105. Published by Munksgaard, 1981 [1]
  31. ^ a b c Arthur C. Guyton, John E. Hall Tratado de fisiología médica (in Spanish). Published by Elsevier España, 2007; page 76. ISBN 84-8174-926-5
  32. ^ a b Simons (2004). "Selective contribution of eukaryotic prefoldin subunits to actin and tubulin binding". J Biol Chem. 279 (6): 4196–203. doi:10.1074/jbc.M306053200. PMID 14634002. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help); Unknown parameter |name= ignored (help)CS1 maint: unflagged free DOI (link)
  33. ^ Martín-Benito (2002). "Structure of eukaryotic prefoldin and of its complexes with unfolded actin and the cytosolic chaperonin CCT". EMBO J. 21 (23): 6377–86. doi:10.1093/emboj/cdf640. PMID 12456645. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help)
  34. ^ a b Vandamme (2009). "alpha-Skeletal muscle actin nemaline myopathy mutants cause cell death in cultured muscle cells". Biochim Biophys Acta. 1793 (7): 1259–71. PMID 19393268. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help); Unknown parameter |name= ignored (help)
  35. ^ a b Brackley (2009). "Activities of the chaperonin containing TCP-1 (CCT): implications for cell cycle progression and cytoskeletal organisation". Cell Stress Chaperones. 14 (1): 23–31. doi:10.1007/s12192-008-0057-x. PMID 18595008. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help)
  36. ^ a b Stirling (2006). "PhLP3 modulates CCT-mediated actin and tubulin folding via ternary complexes with substrates". J Biol Chem. 281 (11): 7012–21. doi:10.1074/jbc.M513235200. PMID 16415341. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help); Unknown parameter |name= ignored (help)CS1 maint: unflagged free DOI (link)
  37. ^ Hansen (1999). "Prefoldin-nascent chain complexes in the folding of cytoskeletal proteins". J Cell Biol. (145). 265-7: 2. PMID 10209023. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help); Unknown parameter |name= ignored (help)
  38. ^ Neirynck (2006). "Actin interacts with CCT via discrete binding sites: a binding transition-release model for CCT-mediated actin folding". J Mol Biol. 355 (1): 124–38. PMID 16300788. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help)
  39. ^ a b c Vavylonis (2005). "Actin polymerization kinetics, cap structure, and fluctuations". Proc Natl Acad Sci U S A. 102 (24): 8543–8. doi:10.1073/pnas.0501435102. PMID 15939882. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help); Unknown parameter |name= ignored (help)
  40. ^ Vandekerckhove J, Weber K (1978). "At least six different actins are expressed in a higher mammal: an analysis based on the amino acid sequence of the amino-terminal tryptic peptide". J. Mol. Biol. 126 (4): 783–802. doi:10.1016/0022-2836(78)90020-7. PMID 745245. {{cite journal}}: Unknown parameter |month= ignored (help)
  41. ^ Khaitlina SY (2001). "Functional specificity of actin isoforms". Int. Rev. Cytol. 202: 35–98. doi:10.1016/S0074-7696(01)02003-4. PMID 11061563.
  42. ^ Garner EC, Campbell CS, Weibel DB, Mullins RD (2007). "Reconstitution of DNA segregation driven by assembly of a prokaryotic actin homolog". Science. 315 (5816): 1270–4. doi:10.1126/science.1138527. PMC 2851738. PMID 17332412. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  43. ^ Kawamura; Maruyama (1970). "Electron Microscopic Particle Length of F-Actin Polymerized in Vitro". 67 (3): 437. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |pub-journal= ignored (help)
  44. ^ "Chapter 12: The Cytoskeleton and Cell Movement". The cell: a molecular approach. ASM Press, Washington. 2007. ISBN 0878932194. {{cite book}}: Cite has empty unknown parameter: |1= (help); Unknown parameter |co-authors= ignored (help); Unknown parameter |name= ignored (help); Unknown parameter |surnames= ignored (help)
  45. ^ "Table of the [[equilibrium constants|constants]] of association and dissociation for the different types of monomers to the actin filament, according to the scientific literature". {{cite web}}: URL–wikilink conflict (help)
  46. ^ Kirschner (1980). "Implications of treadmilling for the stability and polarity of actin and tubulin polymers in vivo" (w). 86 (1): 330–334. PMID 6893454. {{cite journal}}: Cite journal requires |journal= (help); Text "10.1083/jcb.86.1.330" ignored (help)
  47. ^ Cells (Google books). Jones & Bartlett Publishers. 2006. ISBN 9780763739058. {{cite book}}: Unknown parameter |name= ignored (help); Unknown parameter |surnames= ignored (help)
  48. ^ Zhang (15-01-2012). "Multi-isotope imaging mass spectrometry reveals slow protein turnover in hair-cell stereocilia". Nature. 481 (7382): 520–524. doi:10.1038/nature10745. ISSN 1476-4687 0028-0836, 1476-4687. {{cite journal}}: Check |issn= value (help); Check date values in: |date= (help); Unknown parameter |access date= ignored (|access-date= suggested) (help); Unknown parameter |co-authors= ignored (help)
  49. ^ THE STRUCTURE OF CRYSTALLINE PROFILIN-BETA-ACTIN Protein Data Bank
  50. ^ Domínguez (2004). "Actin-binding proteins--a unifying hypothesis". 29 (11): 572–8. PMID 15501675. {{cite journal}}: Cite journal requires |journal= (help); Text "Trends Biochem Sci" ignored (help)
  51. ^ Goldschmidt-clermont; Furman; Wachsstock; Safer; Nachmias; Pollard (1992). "The control of actin nucleotide exchange by thymosin β-4 and profilin. A potential regulatory mechanism for actin polymerization in cells". 3 (9): 1015–1024. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |name4= ignored (help); Unknown parameter |name5= ignored (help); Unknown parameter |name6= ignored (help); Unknown parameter |pub-journal= ignored (help)
  52. ^ Witke, W., Podtelejnikov, A., Di Nardo, A., Sutherland, J., Gurniak, C., Dotti, C., and M. Mann (1998) In Mouse Brain Profilin I and Profilin II Associate With Regulators of the Endocytic Pathway and Actin Assembly. The EMBO Journal 17(4): 967-976 Template:Entrez Pubmed
  53. ^ Carlsson L, Nyström LE, Sundkvist I, Markey F, Lindberg U. (1977) Actin polymerizability is influenced by profilin, a low molecular weight protein in non-muscle cells. J. Mol. Biol. 115:465-483 Template:Entrez Pubmed
  54. ^ Kiselar, J., Janmey, P., Almo, S., Chance, M. (2003). "Visualizing the Ca2+-dependent activation of gelsolin by using synchrotron footprinting". PNAS. 100 (7): 3942–3947. doi:10.1073/pnas.0736004100. PMID 12655044.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  55. ^ Southwick (2000). "Gelsolin and ADF/cofilin enhance the actin dynamics of motile cells". 97 (13): 6936. doi:10.1073/pnas.97.13.6936. PMID 10860951. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name= ignored (help); Unknown parameter |pub-journal= ignored (help)
  56. ^ Caldwell; Heiss; Mermall; Cooper (1989). "Effects of CapZ, an actin-capping protein of muscle, on the polymerization of actin" (w). 28 (21): 8506–8514. doi:10.1021/bi00447a036. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |name4= ignored (help); Unknown parameter |pub-journal= ignored (help)
  57. ^ Weber; Pennise; Babcock; Fowler (1994). "Tropomodulin caps the pointed ends of actin filaments" (PDF). 127 (6): 1627–1635. doi:10.1083/jcb.127.6.1627. PMID 7798317. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |name4= ignored (help); Unknown parameter |pub-journal= ignored (help)
  58. ^ Robinson RC, Turbedsky K, Kaiser DA, Marchand JB, Higgs HN, Choe S, Pollard TD. (2001) Crystal structure of Arp2/3 complex. Science 294(5547):1679-84.
  59. ^ Mullins, R. D. "Structure and function of the Arp2/3 complex". 9 (2). Elsevier: 244–249. doi:10.1016/S0959-440X(99)80034-7. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |co-authors= ignored (help); Unknown parameter |fecha= ignored (|date= suggested) (help); Unknown parameter |fechaaceso= ignored (help); Unknown parameter |revista= ignored (help)
  60. ^ Laura M Machesky (February 1999). "The Arp2/3 complex: a multifunctional actin organizer". 11 (1): 117–121. doi:doi:10.1016/S0955-0674(99)80014-3. {{cite journal}}: Check |doi= value (help); Cite journal requires |journal= (help); Unknown parameter |coauthors= ignored (|author= suggested) (help); Unknown parameter |revista= ignored (help)
  61. ^ Morton; Ayscough; McLaughlin (2000). "Latrunculin alters the actin-monomer subunit interface to prevent polymerization" (PDF). 2 (6): 376–378. doi:10.1038/35014075. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |pub-journal= ignored (help)
  62. ^ a b Cooper (1987). "Effects of cytochalasin and phalloidin on actin" (PDF). 105 (4): 1473–1478. doi:10.1083/jcb.105.4.1473. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name= ignored (help); Unknown parameter |pub-journal= ignored (help)
  63. ^ Rubtsova; Kondratov; Kopnin; Chumakov; Kopnin; Vasiliev (1998). "Disruption of actin microfilaments by cytochalasin D leads to activation of p53". 430 (3): 353–357. doi:10.1016/S0014-5793(98)00692-9. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |name4= ignored (help); Unknown parameter |name5= ignored (help); Unknown parameter |name6= ignored (help); Unknown parameter |pub-journal= ignored (help)
  64. ^ Staves; Wayne; Leopold (1997). "Cytochalasin D does not inhibit gravitropism in roots" (PDF). 84 (11): 1530–1530. doi:10.2307/2446614. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |pub-journal= ignored (help)
  65. ^ a b Grummt (2006). "Actin and myosin as transcription factors" (w). 16 (2): 191–196. doi:10.1016/j.gde.2006.02.001. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name= ignored (help); Unknown parameter |pub-journal= ignored (help)
  66. ^ M.V. Wakelam (2005). "Phospholipase D activity is essential for actin localization and actin-based motility in Dictyostelium" (w). 389 (Pt 1): 207. doi:10.1042/BJ20050085. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name= ignored (help); Unknown parameter |pub-journal= ignored (help)
  67. ^ Eckert Fisiología animal (4ª ed.). ISBN 84-486-0200-5. {{cite book}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |name= ignored (help); Unknown parameter |surnames= ignored (help)
  68. ^ Trombocitopenias (in Spanish) (2nd ed.). Elsevier, España. 2001. p. 25. ISBN 8481745952. {{cite book}}: Unknown parameter |surnames= ignored (help)
  69. ^ a b c Paniagua, R.; Nistal, M.; Sesma, P.; Álvarez-Uría, M.; Fraile, B.; Anadón, R. and José Sáez, F. (2002). Citología e histología vegetal y animal. McGraw-Hill Interamericana de España, S.A.U. ISBN 84-486-0436-9. {{cite book}}: Unknown parameter |Language= ignored (|language= suggested) (help)CS1 maint: multiple names: authors list (link)
  70. ^ a b Moseley; Goode (2006). "The Yeast Actin Cytoskeleton: from Cellular Function to Biochemical Mechanism". 70 (3): 605–645. doi:10.1128/MMBR.00013-06. PMID 16959963. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |pub-journal= ignored (help)
  71. ^ Meagher; McKinney; Kandasamy (1999). "Isovariant Dynamics Expand and Buffer the Responses of Complex Systems: The Diverse Plant Actin Gene Family". 11 (6): 995–1006. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |pub-journal= ignored (help)
  72. ^ Entrada PDBe para 1unc. EBI.
  73. ^ a b Higaki; Sano; Hasezawa (2007). "Actin microfilament dynamics and actin side-binding proteins in plants" (PDF). 10 (6): 549–556. doi:10.1016/j.pbi.2007.08.012. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |pub-journal= ignored (help)
  74. ^ Kovar; Staiger; Weaver; McCurdy (2000). "AtFim1 is an actin filament crosslinking protein from Arabidopsis thaliana". 24 (5): 625–636. doi:10.1046/j.1365-313x.2000.00907.x. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |name4= ignored (help); Unknown parameter |pub-journal= ignored (help)
  75. ^ Zheng B, Han M, Bernier M, Wen JK (2009). "Nuclear actin and actin-binding proteins in the regulation of transcription and gene expression". FEBS J. 276 (10): 2669–85. doi:10.1111/j.1742-4658.2009.06986.x. PMC 2978034. PMID 19459931. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  76. ^ Antoni Bayés de Luna, VV Staff, José López-Sendón, Fause Attie, Eduardo Alegría Ezquerra Cardiología Clínica (in Spanish). Published by Elsevier España, 2002; page 19. ISBN 84-458-1179-7
  77. ^ a b John W. Baynes, Marek H. Dominiczak Bioquímica médica Published by Elsevier Spain, 2007; page 268. ISBN 84-8174-866-8
  78. ^ Eckert Fisiología animal (4ª ed.). ISBN 84-486-0200-5. {{cite book}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |name= ignored (help); Unknown parameter |surnames= ignored (help)
  79. ^ Fujiwara; Porter; Pollard (1978). "Alpha-actinin localization in the cleavage furrow during cytokinesis" (PDF). 79 (1): 268–275. doi:10.1083/jcb.79.1.268. PMID 359574. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |pub-journal= ignored (help)
  80. ^ Pelham; Chang (2002). "Actin dynamics in the contractile ring during cytokinesis in fission yeast". 419 (6902): 82–86. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |pub-journal= ignored (help)
  81. ^ Mashima; Naito; Noguchi; Miller; Nicholson; Tsuruo (1997). "Actin cleavage by CPP-32/apopain during the development of apoptosis" (w). 14 (9): 1007–1012. doi:10.1038/sj.onc.1200919. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |name4= ignored (help); Unknown parameter |name5= ignored (help); Unknown parameter |name6= ignored (help); Unknown parameter |pub-journal= ignored (help)
  82. ^ Wang (2000). "Calpain and caspase: can you tell the difference?" (w). 23 (1): 20–26. doi:10.1016/S0166-2236(99)01479-4. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name= ignored (help); Unknown parameter |pub-journal= ignored (help)
  83. ^ Villa (1998). "Calpain inhibitors, but not caspase inhibitors, prevent actin proteolysis and DNA fragmentation during apoptosis" (PDF). 111 (6): 713–722. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name= ignored (help); Unknown parameter |pub-journal= ignored (help)
  84. ^ Huot; Houle; Rousseau; Deschesnes; Shah; Landry (1998). "SAPK2/p38-dependent F-Actin Reorganization Regulates Early Membrane Blebbing during Stress-induced Apoptosis". 143 (5): 1361–1373. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |name4= ignored (help); Unknown parameter |name5= ignored (help); Unknown parameter |name6= ignored (help); Unknown parameter |pub-journal= ignored (help)
  85. ^ Adams; Nelson; Smith (1996). "Quantitative analysis of cadherin-catenin-actin reorganization during development of cell-cell adhesion" (w). 135 (6): 1899–1911. doi:10.1083/jcb.135.6.1899. PMID 8991100. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |pub-journal= ignored (help)
  86. ^ Witke; Schleicher; Noegel (1992). "Redundancy in the microfilament system: abnormal development of Dictyostelium cells lacking two F-actin cross-linking proteins". 68 (1): 53–62. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |pub-journal= ignored (help)
  87. ^ Fernandez-valle; Gorman; Gomez; Bunge (1997). "Actin Plays a Role in Both Changes in Cell Shape and Gene- Expression Associated with Schwann Cell Myelination". 17 (1): 241–250. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |name4= ignored (help); Unknown parameter |pub-journal= ignored (help)
  88. ^ Wolyniak; Sundstrom (2007). "Role of Actin Cytoskeletal Dynamics in Activation of the Cyclic AMP Pathway and HWP1 Gene Expression in Candida albicans". 6 (10): 1824–1840. doi:10.1128/EC.00188-07. PMID 17715368. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |pub-journal= ignored (help)
  89. ^ Tanaka; Iguchi; Egydio De Carvalho; Tadokoro; Yomogida; Nishimune (2003). "Novel Actin-Like Proteins T-ACTIN 1 and T-ACTIN 2 Are Differentially Expressed in the Cytoplasm and Nucleus of Mouse Haploid Germ Cells" (PDF). 69 (2): 475–482. doi:10.1095/biolreprod.103.015867. PMID 12672658. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |name4= ignored (help); Unknown parameter |name5= ignored (help); Unknown parameter |name6= ignored (help); Unknown parameter |pub-journal= ignored (help)
  90. ^ Jiang; Stillman (1996). "Epigenetic effects on yeast transcription caused by mutations in an actin-related protein present in the nucleus" (w). 10 (5): 604–619. doi:10.1101/gad.10.5.604. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |pub-journal= ignored (help)
  91. ^ Manor; Kachar (2008). "Dynamic length regulation of sensory stereocilia". 19 (6): 502–510. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |access date= ignored (|access-date= suggested) (help); Unknown parameter |name2= ignored (help)
  92. ^ Rzadzinska; Schneider; Davies; Riordan; Kachar (2004). "An actin molecular treadmill and myosins maintain stereocilia functional architecture and self-". 164 (6): 887–897. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |access date= ignored (|access-date= suggested) (help); Unknown parameter |name2= ignored (help)
  93. ^ Su AI, Wiltshire T, Batalov S, Lapp H, Ching KA, Block D, Zhang J, Soden R, Hayakawa M, Kreiman G, Cooke MP, Walker JR, Hogenesch JB (2004). "A gene atlas of the mouse and human protein-encoding transcriptomes". Proc Natl Acad Sci U S A. 101 (16). PMID 15075390.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  94. ^ a b "Uniprot ACTA1". {{cite web}}: Unknown parameter |access date= ignored (|access-date= suggested) (help); Unknown parameter |access year= ignored (help)
  95. ^ a b Friederike S Bathe, Heidi Rommelaere y Laura M Machesky (2007): Phenotypes of Myopathy-related Actin Mutants in differentiated C2C12 Myotubes. BMC Cell Biology, 8:2. doi:10.1186/1471-2121-8-2
  96. ^ Kaindl AM, Rüschendorf F, Krause S, Goebel HH, Koehler K, Becker C, Pongratz D, Müller-Höcker J, Nürnberg P, Stoltenburg-Didinger G, Lochmüller H, Huebner A. (2004). "Missense mutations of ACTA1 cause dominant congenital myopathy with cores". J Med Genet. 41 (11). PMID 15520409.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  97. ^ Sparrow (2003). "Muscle disease caused by mutations in the skeletal muscle alpha-actin gene (ACTA1)". Neuromuscul Disord. 13 (8): 519–31. PMID 12921789. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help); Unknown parameter |name= ignored (help)
  98. ^ North, Kathryn (2002). "Nemaline Myopathy". Gene Reviews. PMID.
  99. ^ Ilkovski B, Nowak KJ, Domazetovska A, Maxwell AL, Clement S, Davies KE, Laing NG, North KN, Cooper ST. (2004). "Evidence for a dominant-negative effect in ACTA1 nemaline myopathy caused by abnormal folding, aggregation and altered polymerization of mutant actin isoforms". Hum. Mol. Genet. 13 (16). PMID 15198992.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  100. ^ Clarke NF, Ilkovski B, Cooper S, Valova VA, Robinson PJ, Nonaka I, Feng JJ, Marston S, North K (2007). "The pathogenesis of ACTA1-related congenital fibre type disproportion". Ann Neurol. 61 (6). PMID 17387733.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  101. ^ "Structure, chromosome location, and expression of the human smooth muscle (enteric type) γ-actin gene: evolution of six human actin genes". Mol Cell Biol. 11 (6). 1991. PMID 1710027.
  102. ^ Watson MB, Lind MJ, Smith L, Drew PJ, Cawkwell L. (2007). "Expression microarray analysis reveals genes associated with in vitro resistance to cisplatin in a cell line model". Acta Oncol. 46 (5). PMID 17562441.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  103. ^ Guo DC, Pannu H, Tran-Fadulu V, Papke CL, Yu RK, Avidan N, Bourgeois S, Estrera AL, Safi HJ, Sparks E, Amor D, Ades L, McConnell V, Willoughby CE, Abuelo D, Willing M, Lewis RA, Kim DH, Scherer S, Tung PP, Ahn C, Buja LM, Raman CS, Shete SS, Milewicz DM. (2007). "Mutations in smooth muscle alpha-actin (ACTA2) lead to thoracic aortic aneurysms and dissections". Nature Genet. 39 (12). PMID 17994018.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  104. ^ Guo (2009). "Mutations in smooth muscle alpha-actin (ACTA2) cause coronary artery disease, stroke, and Moyamoya disease, along with thoracic aortic disease". Am J Human Genet. 84 (5): 617–27. PMID 19409525. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help)
  105. ^ Akpolat N, Yahsi S, Godekmerdan A, Yalniz M, Demirbag K (2005). "The value of alpha-SMA in the evaluation of hepatic fibrosis severity in hepatitis B infection and cirrhosis development: a histopathological and immunohistochemical study". Histopathology. 47 (3). PMID 16115228.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  106. ^ Hamada H, Petrino MG, Kakunaga T. (1982). "Molecular structure and evolutionary origin of human cardiac muscle actin gene". Proc Natl Acad Sci U S A. 79 (19). PMID 6310553.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  107. ^ a b Olson TM, Michels VV, Thibodeau SN, Tai YS, Keating MT. (1998). "Actin mutations in dilated cardiomyopathy, a heritable form of heart failure". Science. 280 (5364). PMID 9563954.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  108. ^ Xia X-G, Zhou H, Samper E, Melov S, Xu Z (2006) Pol II–Expressed shRNA Knocks Down Sod2 Gene Expression and Causes Phenotypes of the Gene Knockout in Mice. PLoS Genet 2(1): e10. doi:10.1371/journal.pgen.0020010
  109. ^ "OMIM 102540". {{cite web}}: Unknown parameter |access date= ignored (|access-date= suggested) (help); Unknown parameter |access year= ignored (help)
  110. ^ Matsson H, Eason J, Bookwalter CS, Klar J, Gustavsson P, Sunnegårdh J, Enell H, Jonzon A, Vikkula M, Gutiérrez I, Granados-Riveron J, Pope M, Bu'Lock F, Cox J, Robinson TE, Song F, Brook DJ, Marston S, Trybus KM, Dahl N. (2008). "Alpha-cardiac actin mutations produce atrial septal defects". Hum Mol Genet. 17 (2). PMID 17947298.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  111. ^ Kabaeva, Kabaeva,Z (2003). "Genetic analysis in hypertrophic cardiomyopathy: missense mutations in the ventricular myosin regulatory light chain gene" (Dissertation). University of Humboldt Berlin. {{cite web}}: External link in |author link= (help); Unknown parameter |access date= ignored (|access-date= suggested) (help)CS1 maint: multiple names: authors list (link)
  112. ^ Olson TM, Doan TP, Kishimoto NY, Whitby FG, Ackerman MJ, Fananapazir L. Inherited and de novo mutations in the cardiac actin gene cause hypertrophic cardiomyopathy. J Mol Cell Cardiol 2000; 32:1687-1694.
  113. ^ "Cardiomiopatía hipertrófica familiar: Genes, mutaciones y modelos animales. Revisión" (in Spanish). Invest. clín, mar. 2004, vol.45, no.1, p.69-100. 2004. {{cite web}}: Unknown parameter |access date= ignored (|access-date= suggested) (help); Unknown parameter |last name= ignored (help)
  114. ^ Kaski JP, Syrris P, Burch M, Tomé-Esteban MT, Fenton M, Christiansen M, Andersen PS, Sebire N, Ashworth M, Deanfield JE, McKenna WJ, Elliott PM. (2008). "Idiopathic restrictive cardiomyopathy in children is caused by mutations in cardiac sarcomere protein genes". Heart. 94 (11). PMID 18467357.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  115. ^ Klaassen S, Probst S, Oechslin E, Gerull B, Krings G, Schuler P, Greutmann M, Hürlimann D, Yegitbasi M, Pons L, Gramlich M, Drenckhahn JD, Heuser A, Berger F, Jenni R, Thierfelder L. (2008). "Mutations in sarcomere protein genes in left ventricular noncompaction". Circulation. 117 (22). PMID 1850600.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  116. ^ "ACEVieW:Homo sapiens complex locus ACTB, encoding actin, β." {{cite web}}: Unknown parameter |access date= ignored (|access-date= suggested) (help); Unknown parameter |access year= ignored (help)
  117. ^ Chang KW, Yang PY, Lai HY, Yeh TS, Chen TC, Yeh CT. (2006). "Identification of a novel actin isoform in hepatocellular carcinoma". Hepatol Res. 36 (1). PMID 16824795.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  118. ^ Kristy L Williams, Masuma Rahimtula and Karen M Mearow. Hsp27 and axonal growth in adult sensory neurons in vitro BMC Neuroscience 2005, 6:24doi:10.1186/1471-2202-6-24
  119. ^ "Pericytoma with t(7;12)". Atlas of Genetics and Cytogenetics in Oncology and Haematology. {{cite web}}: Unknown parameter |access date= ignored (|access-date= suggested) (help)
  120. ^ Procaccio V, Salazar G, Ono S, Styers ML, Gearing M, Dávila A, Jiménez R, Juncos J, Gutekunst CA, Meroni G, Fontanella B, Sontag E, Sontag JM, Faundez V, Wainer BH. (2006). "A mutation of β -actin that alters depolymerization dynamics is associated with autosomal dominant developmental malformations, deafness, and dystonia". Am J Hum Genet. 78 (6). PMID 16685646.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  121. ^ Nunoi H, Yamazaki T, Tsuchiya H, Kato S, Malech HL, Matsuda I, Kanegasaki S. (1999). "A heterozygous mutation of β-actin associated with neutrophil dysfunction and recurrent infection". Proc Natl Acad Sci U S A. 96 (15). PMID 10411937.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  122. ^ a b "ACEVIEW ACTG1". {{cite web}}: Unknown parameter |access date= ignored (|access-date= suggested) (help); Unknown parameter |access year= ignored (help)
  123. ^ Erba HP, Gunning P, Kedes L. (1986). "Nucleotide sequence of the human γ-cytoskeletal actin mRNA: anomalous evolution of vertebrate non-muscle actin genes". Nucleic Acids Res. 14 (13). PMID 373740.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  124. ^ Bryan (2009). "Allele-specific Effects of Human Deafness {γ}-Actin Mutations (DFNA20/26) on the Actin/Cofilin Interaction". J Biol Chem. 284 (27): 18260–9. PMID 19419963. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help); Unknown parameter |name= ignored (help)
  125. ^ Sonnemann KJ, Fitzsimons DP, Patel JR, Liu Y, Schneider MF, Moss RL, Ervasti JM. (2006). "Cytoplasmic γ-actin is not required for skeletal muscle development but its absence leads to a progressive myopathy". Dev Cell. 11 (3). PMID 16950128.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  126. ^ Gouin (1999). "A comparative study of the actin-based motilities of the pathogenic bacteria Listeria monocytogenes, Shigella flexneri and Rickettsia conorii" (w). 112 (11): 1697–1708. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name= ignored (help); Unknown parameter |pub-journal= ignored (help)
  127. ^ Lambrechts (2008). "Listeria comet tails: the actin-based motility machinery at work". Trends Cell Biol. 18 (5): 220–7. PMID 18396046. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help); Unknown parameter |name= ignored (help)
  128. ^ Gouin; Welch; Cossart (2005). "Actin-based motility of intracellular pathogens". 8 (1): 35–45. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |pub-journal= ignored (help)
  129. ^ Parks (2009). "Neutrophil enhancement of Pseudomonas aeruginosa biofilm development: human F-actin and DNA as targets for therapy". J med microbiol: 492-502 pmid=19273646. doi:10.1099/jmm.0.005728-0. {{cite journal}}: Missing pipe in: |pages= (help); Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help); Unknown parameter |name= ignored (help)
  130. ^ Liu (2009). "HIV infection of T cells: actin-in and actin-out". Sci Signal. 2. PMID 19366992. {{cite journal}}: Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help)
  131. ^ Machesky LM (2009). "Actin-Based Protrusions: Promoters or Inhibitors of Cancer Invasion?". Cancer Cell. 16 (1): 5–7. PMID 19573806. {{cite journal}}: Cite has empty unknown parameters: |1= and |2= (help); Unknown parameter |co-authors= ignored (help); Unknown parameter |month= ignored (help)
  132. ^ Miwa (1991). "The nucleotide sequence of a human smooth muscle (enteric type) γ-actin cDNA". 11 (6): 3296–3306. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name= ignored (help); Unknown parameter |pub-journal= ignored (help)
  133. ^ Erickson (2007). "Evolution of the cytoskeleton". 29 (7): 668. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name= ignored (help); Unknown parameter |pub-journal= ignored (help)
  134. ^ Gardiner (2008). "Are histones, tubulin, and actin derived from a common ancestral protein?". 233: 1. doi:10.1007/s00709-008-0305-z. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name= ignored (help); Unknown parameter |pub-journal= ignored (help)
  135. ^ J.A (2009). "Actin and endocytosis: mechanisms and phylogeny". 21: 20. doi:10.1016/j.ceb.2009.01.0. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |pub-journal= ignored (help)
  136. ^ Van Den Ent; Amos; Löwe (2001). "Prokaryotic origin of the actin cytoskeleton" (w). 413: 39–44. doi:10.1038/35092500. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name1= ignored (help); Unknown parameter |name2= ignored (help); Unknown parameter |name3= ignored (help); Unknown parameter |pub-journal= ignored (help)
  137. ^ Carballido-lopez (2006). "The Bacterial Actin-Like Cytoskeleton" (w). 70 (4): 888–909. doi:10.1128/MMBR.00014-06. PMID 17158703. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name= ignored (help); Unknown parameter |pub-journal= ignored (help)
  138. ^ J (2001). "Light-controlled molecular shuttles made from motor proteins carrying cargo on engineered surfaces". 1 (5): 235–239. doi:10.1021/nl015521e. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |pub-journal= ignored (help)
  139. ^ Mansson (2005). "Actin-Based Molecular Motors for Cargo Transportation in Nanotechnology—Potentials and Challenges". 28 (4): 547–555. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |name= ignored (help); Unknown parameter |pub-journal= ignored (help)
  140. ^ Vandesompele J, De Preter K, Pattyn F, Poppe B, Van Roy N, De Paepe A, Speleman F (2002) Accurate normalisation of real-time quantitative RT-PCR data by geometric averaging of multiple internal control genes. Gen. Biol. 3: 1–12.
  141. ^ Selvey; Thompson; Matthaei; Lea; Irving; Griffiths (2001). "ß-Actin". 15 (5): 307–311. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |access date= ignored (|access-date= suggested) (help); Unknown parameter |name2= ignored (help)
  142. ^ Mukai; Schollmeyer; Rosai (1981). "Immunohistochemical localization of actin: Applications in surgical pathology". 5 (1): 91. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |access date= ignored (|access-date= suggested) (help); Unknown parameter |name2= ignored (help)
  143. ^ Haddad; Roy; Zhong; Edgerton; Baldwin (2003). "Atrophy responses to muscle inactivity. II. Molecular markers of protein deficits". 95 (2): 791–802. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |access date= ignored (|access-date= suggested) (help); Unknown parameter |name2= ignored (help)
  144. ^ Remignon; Molette; Babile; Fernandez (2006). "Current advances in proteomic analysis and its use for the resolution of poultry meat quality" (PDF). 62 (01): 123–130. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |access date= ignored (|access-date= suggested) (help); Unknown parameter |name2= ignored (help); Unknown parameter |name4= ignored (help)
  145. ^ M.L. (2004). "Methods and Instruments in Applied Food Analysis". {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |access date= ignored (|access-date= suggested) (help)

Referencias

Template:Link FA

External links


Cite error: There are <ref group=n.> tags on this page, but the references will not show without a {{reflist|group=n.}} template (see the help page).